Skip to main content Accessibility help
×
Hostname: page-component-7479d7b7d-wxhwt Total loading time: 0 Render date: 2024-07-09T06:46:54.391Z Has data issue: false hasContentIssue false

Systemic Psychophysiology

Published online by Cambridge University Press:  27 January 2017

John T. Cacioppo
Affiliation:
University of Chicago
Louis G. Tassinary
Affiliation:
Texas A & M University
Gary G. Berntson
Affiliation:
Ohio State University
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2016

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

References

Aguirre, G. K., Zarahn, E., & D’Esposito, M. (1998). The variability of human, BOLD hemodynamic responses. NeuroImage, 8: 360369.Google Scholar
Amunts, K., Schleicher, A., & Zilles, K. (2007). Cytoarchitecture of the cerebral cortex: more than localization. NeuroImage, 37: 10611065.CrossRefGoogle ScholarPubMed
Andersson, J. L., Hutton, C., Ashburner, J., Turner, R., & Friston, K. (2001). Modeling geometric deformations in EPI time series. NeuroImage, 13: 903919.Google Scholar
Andrews-Hanna, J. R., Reidler, J. S., Huang, C., & Buckner, R. L. (2010). Evidence for the default network’s role in spontaneous cognition. Journal of Neurophysiology, 104: 322335.Google Scholar
Aron, A., Fisher, H., Mashek, D. J., Strong, G., Li, H., & Brown, L. L. (2005). Reward, motivation, and emotion systems associated with early-stage intense romantic love. Journal of Neurophysiology, 94: 327337.Google Scholar
Ashburner, J. (2007). A fast diffeomorphic image registration algorithm. NeuroImage, 38: 95113.CrossRefGoogle ScholarPubMed
Ashburner, J. & Friston, K. J. (2000). Voxel-based morphometry: the methods. NeuroImage, 11: 805821.CrossRefGoogle ScholarPubMed
Atlas, L. Y., Lindquist, M. A., Bolger, N., & Wager, T. D. (2014). Brain mediators of the effects of noxious heat on pain. Pain, 155: 16321648.Google Scholar
Atlas, L. Y., Whittington, R. A., Lindquist, M. A., Wielgosz, J., Sonty, N., & Wager, T. D. (2012). Dissociable influences of opiates and expectations on pain. Journal of Neuroscience, 32: 80538064.CrossRefGoogle ScholarPubMed
Bates, D., Maechler, M., Bolker, B., & Walker, S. (2013). LME4: Linear mixed-effects models using Eigen and S4. R package version 1.Google Scholar
Beckmann, C. F., Jenkinson, M., & Smith, S. M. (2003). General multilevel linear modeling for group analysis in FMRI. NeuroImage, 20: 10521063.Google Scholar
Behrens, T. E. J., Berg, H. J., Jbabdi, S., Rushworth, M. F. S., & Woolrich, M. W. (2007). Probabilistic diffusion tractography with multiple fibre orientations: what can we gain? NeuroImage, 34: 144155.Google Scholar
Bendriem, B. & Townsend, D. W. (1998). The Theory and Practice of 3D PET. Boston and Dordrecht: Kluwer.Google Scholar
Benjamini, Y. & Hochberg, Y. (1995) Controlling the false discovery rate: a practical and powerful approach to multiple testing. Journal of the Royal Statistical Society B, 57: 289300.Google Scholar
Bernstein, M. A., King, K. F., & Zhou, Z. J. (2004). Handbook of MRI Pulse Sequences. Burlington, MA: Elsevier Academic Press.Google Scholar
Binder, J. R., Desai, R. H., Graves, W. W., & Conant, L. L. (2009). Where is the semantic system? A critical review and meta-analysis of 120 functional neuroimaging studies. Cerebral Cortex, 19: 27672796.Google Scholar
Birn, R. M., Saad, Z. S., & Bandettini, P. A. (2001). Spatial heterogeneity of the nonlinear dynamics in the FMRI BOLD response. NeuroImage, 14: 817826.Google Scholar
Biswal, B., Yetkin, F. Z., Haughton, V. M., & Hyde, J. S. (1995). Functional connectivity in the motor cortex of resting human brain using echo-planar MRI. Magnetic Resonance in Medicine, 34: 537541.CrossRefGoogle ScholarPubMed
Bohning, D. E., Pecheny, A. P., Epstein, C. M., Speer, A. M., Vincent, D. J., Dannels, W., & George, M. S. (1997). Mapping transcranial magnetic stimulation (TMS) fields in vivo with MRI. Neuroreport, 8: 25352538.Google Scholar
Bohning, D. E., Shastri, A., McConnell, K. A., Nahas, Z., Lorberbaum, J. P., Roberts, D. R., Teneback, C., Vincent, D. J., & George, M. S. (1999). A combined TMS/fMRI study of intensity-dependent TMS over motor cortex. Biological Psychiatry, 45: 385394.CrossRefGoogle ScholarPubMed
Bornhovd, K., Quante, M., Glauche, V., Bromm, B., Weiller, C., & Buchel, C. (2002). Painful stimuli evoke different stimulus-response functions in the amygdala, prefrontal, insula and somatosensory cortex: a single-trial fMRI study. Brain, 125: 13261336.Google Scholar
Boubela, R. N., Kalcher, K., Huf, W., Seidel, E. M., Derntl, B., Pezawas, L., … & Moser, E. (2015). fMRI measurements of amygdala activation are confounded by stimulus correlated signal fluctuation in nearby veins draining distant brain regions. Scientific Reports, 5: 10499.Google Scholar
Boynton, G. M., Engel, S. A., Glover, G. H., & Heeger, D. J. (1996). Linear systems analysis of functional magnetic resonance imaging in human V1. Journal of Neuroscience, 16: 42074221.Google Scholar
Brett, M., Johnsrude, I. S., & Owen, A. M. (2002). The problem of functional localization in the human brain. Nature Reviews Neuroscience, 3: 243249.CrossRefGoogle ScholarPubMed
Brooks, J. C., Beckmann, C. F., Miller, K. L., Wise, R. G., Porro, C. A., Tracey, I., & Jenkinson, M. (2008). Physiological noise modelling for spinal functional magnetic resonance imaging studies. NeuroImage, 39: 680692.Google Scholar
Brown, A. K., Fujita, M., Fujimura, Y., Liow, J. S., Stabin, M., Ryu, Y. H., … & Innis, R. B. (2007). Radiation dosimetry and biodistribution in monkey and man of 11C-PBR28: a PET radioligand to image inflammation. Journal of Nuclear Medicine, 48: 20722079.Google Scholar
Buchel, C., Bornhovd, K., Quante, M., Glauche, V., Bromm, B., & Weiller, C. (2002), Dissociable neural responses related to pain intensity, stimulus intensity, and stimulus awareness within the anterior cingulate cortex: a parametric single-trial laser functional magnetic resonance imaging study. Journal of Neuroscience, 22: 970976.Google Scholar
Buckner, R. L., Andrews-Hanna, J. R., & Schacter, D. L. (2008). The brain’s default network. Annals of the New York Academy of Sciences, 1124: 138.Google Scholar
Buracas, G. T. & Boynton, G. M. (2002). Efficient design of event-related fMRI experiments using M-sequences. NeuroImage, 16: 801813.Google Scholar
Bush, G., Luu, P., & Posner, M. I. (2000). Cognitive and emotional influences in anterior cingulate cortex. Trends in Cognitive Sciences, 4: 215222.Google Scholar
Button, K. S., Ioannidis, J. P., Mokrysz, C., Nosek, B. A., Flint, J., Robinson, E. S., & Munafo, M. R. (2013). Power failure: why small sample size undermines the reliability of neuroscience. Nature Reviews Neuroscience, 14: 365376.Google Scholar
Buxton, R. B. & Frank, L. R. (1997). A model for the coupling between cerebral blood flow and oxygen metabolism during neural stimulation. Journal of Cerebral Blood Flow & Metabolism, 17: 6472.CrossRefGoogle Scholar
Buxton, R. B., Frank, L. R., Wong, E. C., Siewert, B., Warach, S., & Edelman, R. R. (1998). A general kinetic model for quantitative perfusion imaging with arterial spin labeling. Magnetic Resonance in Medicine, 40: 383396.Google Scholar
Buxton, R. B., Uludag, K., Dubowitz, D. J., & Liu, T. T. (2004). Modeling the hemodynamic response to brain activation. NeuroImage, 23: S220S233.Google Scholar
Cacioppo, J. T., & Tassinary, L. G. (1990). Inferring psychological significance from physiological signals. American Psychologist, 45: 1628.Google Scholar
Calhoun, V. D., Miller, R., Pearlson, G., & Adali, T. (2014). The chronnectome: time-varying connectivity networks as the next frontier in fMRI data discovery. Neuron, 84: 262274.Google Scholar
Chaimow, D., Yacoub, E., Ugurbil, K., & Shmuel, A. (2011). Modeling and analysis of mechanisms underlying fMRI-based decoding of information conveyed in cortical columns. NeuroImage, 56: 627642.Google Scholar
Cheng, K., Waggoner, R. A., & Tanaka, K. (2001). Human ocular dominance columns as revealed by high-field functional magnetic resonance imaging. Neuron, 32: 359374.Google Scholar
Collins, D. L., Neelin, P., Peters, T. M., & Evans, A. C. (1994). Automatic 3D intersubject registration of MR volumetric data in standardized Talairach space. Journal of Computer Assisted Tomography, 18: 192205.Google Scholar
Coltheart, M. (2006). What has functional neuroimaging told us about the mind (so far)? Cortex, 42: 323331.Google Scholar
Constable, R. T. & Spencer, D. D. (1999). Composite image formation in z-shimmed functional MR imaging. Magnetic Resonance in Medicine, 42: 110117.3.0.CO;2-3>CrossRefGoogle ScholarPubMed
Cover, T. M. & Thomas, J. A. (1991). Elements of Information Theory. New York: John Wiley.Google Scholar
Cribben, I., Haraldsdottir, R., Atlas, L. Y., Wager, T. D., & Lindquist, M. A. (2012). Dynamic connectivity regression: determining state-related changes in brain connectivity. NeuroImage, 61: 907920.Google Scholar
de Quervain, D. J., Fischbacher, U., Treyer, V., Schellhammer, M., Schnyder, U., Buck, A., & Fehr, E. (2004). The neural basis of altruistic punishment. Science, 305: 12541258.Google Scholar
Deckers, R. H., van Gelderen, P., Ries, M., Barret, O., Duyn, J. H., Ikonomidou, V. N., … & de Zwart, J. A. (2006). An adaptive filter for suppression of cardiac and respiratory noise in MRI time series data. NeuroImage, 33: 10721081.Google Scholar
Denis Le Bihan, M. D., Mangin, J. F., Poupon, C., Clark, C. A., Pappata, S., Molko, N., & Chabriat, H. (2001). Diffusion tensor imaging: concepts and applications. Journal of Magnetic Resonance Imaging, 13: 534546.Google Scholar
Denny, B. T., Kober, H., Wager, T. D., & Ochsner, K. N. (2012). A meta-analysis of functional neuroimaging studies of self- and other judgments reveals a spatial gradient for mentalizing in medial prefrontal cortex. Journal of Cognitive Neuroscience, 24: 17421752.Google Scholar
Desmond, J. E. & Glover, G. H. (2002). Estimating sample size in functional MRI (fMRI) neuroimaging studies: statistical power analyses. Journal of Neuroscience Methods, 118: 115128.Google Scholar
Detre, J. A., Zhang, W. G., Roberts, D. A., Silva, A. C., Williams, D. S., Grandis, D. J., … & Leigh, J. S. (1994). Tissue-specific perfusion imaging using arterial spin-labeling. NMR in Biomedicine, 7: 7582.Google Scholar
Devlin, J. T. & Poldrack, R. A. (2007). In praise of tedious anatomy. NeuroImage, 37: 10331041.CrossRefGoogle ScholarPubMed
Disbrow, E. A., Slutsky, D. A., Roberts, T. P., & Krubitzer, L. A. (2000). Functional MRI at 1.5 tesla: a comparison of the blood oxygenation level-dependent signal and electrophysiology. Proceedings of the National Academy of Sciences of the USA, 97: 97189723.Google Scholar
Doucet, G., Naveau, M., Petit, L., Zago, L., Crivello, F., Jobard, G., … & Joliot, M. (2012). Patterns of hemodynamic low-frequency oscillations in the brain are modulated by the nature of free thought during rest. NeuroImage, 59: 31943200.Google Scholar
Duong, T. Q., Yacoub, E., Adriany, G., Hu, X., Ugurbil, K., Vaughan, J. T., … & Kim, S. G. (2002). High-resolution, spin-echo BOLD, and CBF fMRI at 4 and 7 T. Magnetic Resonance in Medicine, 48: 589593.Google Scholar
Duvernoy, H. M. (2012). The Human Brain Stem and Cerebellum: Surface, Structure, Vascularization, and Three-Dimensional Sectional Anatomy, with MRI. Dordrecht: Springer Science & Business Media.Google Scholar
Eickhoff, S. B., Stephan, K. E., Mohlberg, H., Grefkes, C., Fink, G. R., Amunts, K., & Zilles, K. (2005). A new SPM toolbox for combining probabilistic cytoarchitectonic maps and functional imaging data. NeuroImage, 25: 13251335.Google Scholar
Eisenberger, N. I., Lieberman, M. D., & Williams, K. D. (2003). Does rejection hurt? An fMRI study of social exclusion. Science, 302: 290292.Google Scholar
Elster, A. D. (1994). Questions and Answers in Magnetic Resonance Imaging. St. Louis, MO: Mosby.Google Scholar
Ethofer, T., Van De Ville, D., Scherer, K., & Vuilleumier, P. (2009). Decoding of emotional information in voice-sensitive cortices. Current Biology, 19: 10281033.Google Scholar
Feinberg, D. A., Moeller, S., Smith, S. M., Auerbach, E., Ramanna, S., Gunther, M., … & Yacoub, E. (2010). Multiplexed echo planar imaging for sub-second whole brain FMRI and fast diffusion imaging. PLoS One, 5: e15710.Google Scholar
Finsterbusch, J., Busch, M. G., & Larson, P. E. Z. (2013). Signal scaling improves the signal-to-noise ratio of measurements with segmented 2D-selective radiofrequency excitations. Magnetic Resonance in Medicine, 70: 14911499.Google Scholar
Fischl, B., Sereno, M. I., & Dale, A. M. (1999). Cortical surface-based analysis. II: Inflation, flattening, and a surface-based coordinate system. NeuroImage, 9: 195207.Google Scholar
Fox, M. D., Snyder, A. Z., Vincent, J. L., Corbetta, M., Van Essen, D. C., & Raichle, M. E. (2005). The human brain is intrinsically organized into dynamic, anticorrelated functional networks. Proceedings of the National Academy of Sciences of the USA, 102: 96739678.Google Scholar
Fox, M. D., Snyder, A. Z., Vincent, J. L., & Raichle, M. E. (2007). Intrinsic fluctuations within cortical systems account for intertrial variability in human behavior. Neuron, 56: 171184.Google Scholar
Frey, K. A. (1999). Positron emission tomography. In Siegel, G. J., Agranoff, B. W., Albers, R. W., Fisher, S. K., & Uhler, M. D. (eds.), Basic Neurochemistry, 6th edn. (pp. 11091131). Philadelphia: Lippincott, Williams, & Wilkins.Google Scholar
Friston, K. J. (2009). Modalities, modes, and models in functional neuroimaging. Science, 326: 399403.Google Scholar
Friston, K. J. (2011). Functional and effective connectivity: a review. Brain Connectivity, 1: 1336.Google Scholar
Friston, K. J. (2012). Ten ironic rules for non-statistical reviewers. NeuroImage, 61: 13001310.Google Scholar
Friston, K. J., Buechel, C., Fink, G. R., Morris, J., Rolls, E., & Dolan, R. J. (1997). Psychophysiological and modulatory interactions in neuroimaging. NeuroImage, 6: 218229.Google Scholar
Friston, K. J., Frith, C. D., Turner, R., & Frackowiak, R. S. (1995). Characterizing evoked hemodynamics with fMRI. NeuroImage, 2: 157165.Google Scholar
Friston, K. J., Harrison, L., & Penny, W. (2003). Dynamic causal modelling. NeuroImage, 19: 12731302.Google Scholar
Friston, K. J., Mechelli, A., Turner, R., & Price, C. J. (2000). Nonlinear responses in fMRI: the Balloon model, Volterra kernels, and other hemodynamics. NeuroImage, 12: 466477.CrossRefGoogle ScholarPubMed
Gianaros, P. J. & Wager, T. D. (2015). Brain–body pathways linking psychological stress and physical health. Current Directions in Psychological Science, 24: 313321.Google Scholar
Glahn, D. C., Paus, T., & Thompson, P. M. (2007a). Imaging genomics: mapping the influence of genetics on brain structure and function. Human Brain Mapping, 28: 461463.Google Scholar
Glahn, D. C., Thompson, P. M., & Blangero, J. (2007b). Neuroimaging endophenotypes: strategies for finding genes influencing brain structure and function. Human Brain Mapping, 28: 488501.Google Scholar
Glover, G. H. & Law, C. S. (2001). Spiral-in/out BOLD fMRI for increased SNR and reduced susceptibility artifacts. Magnetic Resonance in Medicine, 46:515522.Google Scholar
Glover, G. H., Li, T. Q., & Ress, D. (2000). Image-based method for retrospective correction of physiological motion effects in fMRI: RETROICOR. Magnetic Resonance in Medicine, 44: 162167.Google Scholar
Goldman, R. I., Stern, J. M., Engel, J. Jr., & Cohen, M. S. (2000). Acquiring simultaneous EEG and functional MRI. Clinical Neurophysiology, 111: 19741980.Google Scholar
Gonzalez-Castillo, J., Saad, Z. S., Handwerker, D. A., Inati, S. J., Brenowitz, N., & Bandettini, P. A. (2012). Whole-brain, time-locked activation with simple tasks revealed using massive averaging and model-free analysis. Proceedings of the National Academy of Sciences of the USA, 109: 54875492.Google Scholar
Good, C. D., Johnsrude, I. S., Ashburner, J., Henson, R. N. A., Friston, K. J., & Frackowiak, R. S. J. (2001). A voxel-based morphometric study of ageing in 465 normal adult human brains. NeuroImage, 14: 2136.Google Scholar
Grinband, J., Savitskaya, J., Wager, T. D., Teichert, T., Ferrera, V. P., & Hirsch, J. (2011). The dorsal medial frontal cortex is sensitive to time on task, not response conflict or error likelihood. NeuroImage, 57: 303311.Google Scholar
Haacke, E. M. (1999). Magnetic Resonance Imaging: Physical Principles and Sequence Design. New York: John Wiley.Google Scholar
Haines, D. E. (2000). Neuroanatomy: An Atlas of Structures, Sections, and Systems. Philadelphia: Lippincott Williams & Wilkins.Google Scholar
Hare, T. A., Camerer, C. F., & Rangel, A. (2009). Self-control in decision-making involves modulation of the vmPFC valuation system. Science, 324: 646648.Google Scholar
Haxby, J. V., Guntupalli, J. S., Connolly, A. C., Halchenko, Y. O., Conroy, B. R., Gobbini, M. I., … & Ramadge, P. J. (2011). A common, high-dimensional model of the representational space in human ventral temporal cortex. Neuron, 72: 404416.Google Scholar
Haynes, J. D. (2015). A primer on pattern-based approaches to fMRI: principles, pitfalls, and perspectives. Neuron, 87: 257270.Google Scholar
Haynes, J. D., Deichmann, R., & Rees, G. (2005). Eye-specific effects of binocular rivalry in the human lateral geniculate nucleus. Nature, 438: 496499.Google Scholar
Heeger, D. J. & Ress, D. (2002). What does fMRI tell us about neuronal activity? Nature Reviews Neuroscience, 3: 142151.Google Scholar
Henson, R., Shallice, T., & Dolan, R. (2000). Neuroimaging evidence for dissociable forms of repetition priming. Science, 287: 12691272.Google Scholar
Honey, C. J., Sporns, O., Cammoun, L., Gigandet, X., Thiran, J. P., Meuli, R., & Hagmann, P. (2009). Predicting human resting-state functional connectivity from structural connectivity. Proceedings of the National Academy of Sciences of the USA, 106: 20352040.Google Scholar
Horikawa, T., Tamaki, M., Miyawaki, Y., & Kamitani, Y. (2013). Neural decoding of visual imagery during sleep. Science, 340: 639642.Google Scholar
Huettel, S. A., Song, A. W., & McCarthy, G. (2004). Functional Magnetic Resonance Imaging. Sunderland, MA: Sinauer Associates.Google Scholar
Huth, A. G., Nishimoto, S., Vu, A. T., & Gallant, J. L. (2012). A continuous semantic space describes the representation of thousands of object and action categories across the human brain. Neuron, 76: 12101224.Google Scholar
Johansen-Berg, H. & Behrens, T. E. (2006). Just pretty pictures? What diffusion tractography can add in clinical neuroscience. Current Opinion in Neurology, 19: 379385.Google Scholar
Johansen-Berg, H., Behrens, T. E., Robson, M. D., Drobnjak, I., Rushworth, M. F., Brady, J. M., … & Matthews, P. M. (2004). Changes in connectivity profiles define functionally distinct regions in human medial frontal cortex. Proceedings of the National Academy of Sciences of the USA, 101: 1333513340.Google Scholar
Josephs, O. & Henson, R. N. (1999). Event-related functional magnetic resonance imaging: modelling, inference and optimization. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 354: 12151228.Google Scholar
Kamitani, Y. & Tong, F. (2005). Decoding the visual and subjective contents of the human brain. Nature Neuroscience, 8: 679685.Google Scholar
Kao, M. H., Mandal, A., Lazar, N., & Stufken, J. (2009). Multi-objective optimal experimental designs for event-related fMRI studies. NeuroImage, 44: 849856.Google Scholar
Kastner, S. & Ungerleider, L. G. (2000). Mechanisms of visual attention in the human cortex. Annual Review of Neuroscience, 23: 315341.Google Scholar
Kleinschmidt, A., Buchel, C., Zeki, S., & Frackowiak, R. S. (1998). Human brain activity during spontaneously reversing perception of ambiguous figures. Proceedings of the Royal Society of London B: Biological Sciences, 265: 24272433.Google Scholar
Klunk, W. E., Engler, H., Nordberg, A., Wang, Y., Blomqvist, G., Holt, D. P., … & Långström, B. (2004). Imaging brain amyloid in Alzheimer’s disease with Pittsburgh Compound-B. Annals of Neurology, 55: 306319.Google Scholar
Kober, H., Barrett, L. F., Joseph, J., Bliss-Moreau, E., Lindquist, K., & Wager, T. D. (2008). Functional grouping and cortical–subcortical interactions in emotion: a meta-analysis of neuroimaging studies. NeuroImage, 42: 9981031.Google Scholar
Kong, Y., Jenkinson, M., Andersson, J., Tracey, I., & Brooks, J. C. (2012). Assessment of physiological noise modelling methods for functional imaging of the spinal cord. NeuroImage, 60: 15381549.Google Scholar
Kriegeskorte, N., Lindquist, M. A., Nichols, T. E., Poldrack, R. A., & Vul, E. (2010). Everything you never wanted to know about circular analysis, but were afraid to ask. Journal of Cerebral Blood Flow & Metabolism, 30: 15511557.Google Scholar
Kriegeskorte, N., Simmons, W. K., Bellgowan, P. S., & Baker, C. I. (2009). Circular analysis in systems neuroscience: the dangers of double dipping. Nature Neuroscience, 12: 535540.Google Scholar
Kvitsiani, D., Ranade, S., Hangya, B., Taniguchi, H., Huang, J. Z., & Kepecs, A. (2013). Distinct behavioural and network correlates of two interneuron types in prefrontal cortex. Nature, 498: 363366.Google Scholar
Kwong, K. K., Belliveau, J. W., Chesler, D. A., Goldberg, I. E., Weisskoff, R. M., Poncelet, B. P., … & Turner, R. (1992). Dynamic magnetic resonance imaging of human brain activity during primary sensory stimulation. Proceedings of the National Academy of Sciences of the USA, 89: 56755679.Google Scholar
Laufs, H., Daunizeau, J., Carmichael, D. W., & Kleinschmidt, A. (2008). Recent advances in recording electrophysiological data simultaneously with magnetic resonance imaging. NeuroImage, 40: 515528.Google Scholar
Leitao, J., Thielscher, A., Tunnerhoff, J., & Noppeney, U. (2015). Concurrent TMS-fMRI reveals interactions between dorsal and ventral attentional systems. Journal of Neuroscience, 35: 1144511457.Google Scholar
Lindquist, K. A., Wager, T. D., Kober, H., Bliss-Moreau, E., & Barrett, L. F. (2012a). The brain basis of emotion: a meta-analytic review. Behavioral and Brain Sciences, 35: 121143.Google Scholar
Lindquist, M. A., Caffo, B., & Crainiceanu, C. (2013). Ironing out the statistical wrinkles in “ten ironic rules.” NeuroImage, 81: 499502.Google Scholar
Lindquist, M. A., Spicer, J., Asllani, I., & Wager, T. D. (2012b). Estimating and testing variance components in a multi-level GLM. NeuroImage, 59: 490501.Google Scholar
Lindquist, M. A., Waugh, C., & Wager, T. D. (2007). Modeling state-related fMRI activity using change-point theory. NeuroImage, 35: 11251141.Google Scholar
Lindquist, M. A., Zhang, C. H., Glover, G., & Shepp, L. (2008). Rapid three-dimensional functional magnetic resonance imaging of the initial negative BOLD response. Journal of Magnetic Resonance, 191: 100111.Google Scholar
Liu, T. T. (2004). Efficiency, power, and entropy in event-related fMRI with multiple trial types. Part II: design of experiments. NeuroImage, 21: 401413.Google Scholar
Loggia, M. L., Chonde, D. B., Akeju, O., Arabasz, G., Catana, C., Edwards, R. R., … & Hooker, J. M. (2015). Evidence for brain glial activation in chronic pain patients. Brain, 138: 604615.Google Scholar
Logothetis, N. K. (2008). What we can do and what we cannot do with fMRI. Nature, 453: 869878.Google Scholar
Logothetis, N. K., Pauls, J., Augath, M., Trinath, T., & Oeltermann, A. (2001). Neurophysiological investigation of the basis of the fMRI signal. Nature, 412: 150157.Google Scholar
Maguire, E. A., Gadian, D. G., Johnsrude, I. S., Good, C. D., Ashburner, J., Frackowiak, R. S., & Frith, C. D. (2000). Navigation-related structural change in the hippocampi of taxi drivers. Proceedings of the National Academy of Sciences of the USA, 97: 43984403.Google Scholar
Mai, J. K., Paxinos, G., & Voss, T. (2007). Atlas of the Human Brain, 3rd edn. New York: Academic Press.Google Scholar
Menon, R. S. (2002). Postacquisition suppression of large-vessel BOLD signals in high-resolution fMRI. Magnetic Resonance in Medicine, 47: 19.CrossRefGoogle ScholarPubMed
Miezin, F. M., Maccotta, L., Ollinger, J. M., Petersen, S. E., & Buckner, R. L. (2000). Characterizing the hemodynamic response: effects of presentation rate, sampling procedure, and the possibility of ordering brain activity based on relative timing. NeuroImage, 11: 735759.Google Scholar
Morawetz, C., Holz, P., Lange, C., Baudewig, J., Weniger, G., Irle, E., & Dechent, P. (2008). Improved functional mapping of the human amygdala using a standard functional magnetic resonance imaging sequence with simple modifications. Magnetic Resonance Imaging, 26: 4553.Google Scholar
Mumford, J. A. & Nichols, T. E. (2008). Power calculation for group fMRI studies accounting for arbitrary design and temporal autocorrelation. NeuroImage, 39: 261268.Google Scholar
Nichols, T. & Hayasaka, S. (2003). Controlling the familywise error rate in functional neuroimaging: a comparative review. Statistical Methods in Medical Research, 12: 419446.Google Scholar
Nichols, T. E. & Holmes, A. P. (2002). Nonparametric permutation tests for functional neuroimaging: a primer with examples. Human Brain Mapping, 15: 125.Google Scholar
Noll, D. C., Fessler, J. A., & Sutton, B. P. (2005). Conjugate phase MRI reconstruction with spatially variant sample density correction. IEEE Transactions on Medical Imaging, 24: 325336.Google Scholar
Norman, K. A., Polyn, S. M., Detre, G. J., & Haxby, J. V. (2006). Beyond mind-reading: multi-voxel pattern analysis of fMRI data. Trends in Cognitive Sciences, 10: 424430.Google Scholar
Northoff, G., Heinzel, A., de Greck, M., Bermpohl, F., Dobrowolny, H., & Panksepp, J. (2006). Self-referential processing in our brain: a meta-analysis of imaging studies on the self. NeuroImage, 31: 440457.Google Scholar
Ogawa, S., Lee, T. M., Kay, A. R., & Tank, D. W. (1990). Brain magnetic resonance imaging with contrast dependent on blood oxygenation. Proceedings of the National Academy of Sciences of the USA, 87: 98689872.Google Scholar
Ogawa, S., Tank, D. W., Menon, R., Ellermann, J. M., Kim, S. G., Merkle, H., & Ugurbil, K. (1992). Intrinsic signal changes accompanying sensory stimulation: functional brain mapping with magnetic resonance imaging. Proceedings of the National Academy of Sciences of the USA, 89: 59515955.Google Scholar
Paton, J. J., Belova, M. A., Morrison, S. E., & Salzman, C. D. (2006). The primate amygdala represents the positive and negative value of visual stimuli during learning. Nature, 439: 865870.Google Scholar
Paus, T. (2001). Primate anterior cingulate cortex: where motor control, drive and cognition interface. Nature Reviews Neuroscience, 2: 417424.CrossRefGoogle ScholarPubMed
Petrini, K., Pollick, F. E., Dahl, S., McAleer, P., McKay, L. S., Rocchesso, D., … & Puce, A. (2011). Action expertise reduces brain activity for audiovisual matching actions: an fMRI study with expert drummers. NeuroImage, 56: 14801492.Google Scholar
Phillips, C., Rugg, M. D., & Friston, K. J. (2002). Anatomically informed basis functions for EEG source localization: combining functional and anatomical constraints. NeuroImage, 16: 678695.Google Scholar
Poldrack, R. A. (2011). Inferring mental states from neuroimaging data: from reverse inference to large-scale decoding. Neuron, 72: 692697.CrossRefGoogle ScholarPubMed
Price, C. J., Veltman, D. J., Ashburner, J., Josephs, O., & Friston, K. J. (1999). The critical relationship between the timing of stimulus presentation and data acquisition in blocked designs with fMRI. NeuroImage, 10: 3644.CrossRefGoogle ScholarPubMed
Raichle, M. E., MacLeod, A. M., Snyder, A. Z., Powers, W. J., Gusnard, D. A., & Shulman, G. L. (2001). A default mode of brain function. Proceedings of the National Academy of Sciences of the USA, 98: 676682.Google Scholar
Rasbash, J. (2002). A User’s Guide to MLwiN. Centre for Multilevel Modelling, University of London.Google Scholar
Raudenbush, S. W. & Bryk, A. S. (2002). Hierarchical Linear Models: Applications and Data Analysis, 2nd edn. Newbury Park, CA: Sage.Google Scholar
Reiman, E. M., Fusselman, M. J., Fox, P. T., & Raichle, M. E. (1989). Neuroanatomical correlates of anticipatory anxiety. Science, 243: 10711074 [erratum published in Science, 256 (1992): 1696].Google Scholar
Rosen, B. R., Buckner, R. L., & Dale, A. M. (1998). Event-related functional MRI: past, present, and future. Proceedings of the National Academy of Sciences of the USA, 95: 773780.Google Scholar
Ruff, C. C., Blankenburg, F., Bjoertomt, O., Bestmann, S., Freeman, E., Haynes, J. D., … & Driver, J. (2006). Concurrent TMS-fMRI and psychophysics reveal frontal influences on human retinotopic visual cortex. Current Biology, 16: 14791488.Google Scholar
Saad, Z. S., Reynolds, R. C., Argall, B., Japee, S., & Cox, R. W. (2004). SUMA: an interface for surface-based intra- and inter-subject analysis with AFNI. IEEE International Symposium on Biomedical Imaging: Nano to Macro, 1512: 15101513.Google Scholar
Sandler, M. P. (2003). Diagnostic Nuclear Medicine. Philadelphia, PA: Lippincott, Williams & Wilkins.Google Scholar
Sarter, M., Berntson, G. G., & Cacioppo, J. T. (1996). Brain imaging and cognitive neuroscience: toward strong inference in attributing function to structure. American Psychologist, 51: 1321.Google Scholar
Schacter, D. L., Buckner, R. L., Koutstaal, W., Dale, A. M., & Rosen, B. R. (1997). Late onset of anterior prefrontal activity during true and false recognition: an event-related fMRI study. NeuroImage, 6: 259269.Google Scholar
Scheibe, C., Ullsperger, M., Sommer, W., & Heekeren, H. R. (2010). Effects of parametrical and trial-to-trial variation in prior probability processing revealed by simultaneous electroencephalogram/functional magnetic resonance imaging. Journal of Neuroscience, 30: 1670916717.Google Scholar
Seeley, W. W., Menon, V., Schatzberg, A. F., Keller, J., Glover, G. H., Kenna, H., … & Greicius, M. D. (2007). Dissociable intrinsic connectivity networks for salience processing and executive control. Journal of Neuroscience, 27: 23492356.Google Scholar
Setsompop, K., Gagoski, B. A., Polimeni, J. R., Witzel, T., Wedeen, V. J., & Wald, L. L. (2012). Blipped-controlled aliasing in parallel imaging for simultaneous multislice echo planar imaging with reduced g-factor penalty. Magnetic Resonance in Medicine, 67: 12101224.Google Scholar
Shulman, R. G. & Rothman, D. L. (1998). Interpreting functional imaging studies in terms of neurotransmitter cycling. Proceedings of the National Academy of Sciences of the USA, 95: 1199311998.Google Scholar
Shulman, R. G., Rothman, D. L., Behar, K. L., & Hyder, F. (2004). Energetic basis of brain activity: implications for neuroimaging. Trends in Neurosciences, 27: 489495.CrossRefGoogle ScholarPubMed
Sibson, N. R., Dhankhar, A., Mason, G. F., Behar, K. L., Rothman, D. L., & Shulman, R. G. (1997). In vivo 13C NMR measurements of cerebral glutamine synthesis as evidence for glutamate-glutamine cycling. Proceedings of the National Academy of Sciences of the USA, 94: 26992704.Google Scholar
Sinha, R., Lacadie, C., Skudlarski, P., & Wexler, B. E. (2004). Neural circuits underlying emotional distress in humans. Annals of the New York Academy of Sciences, 1032: 254257.Google Scholar
Skudlarski, P., Constable, R. T., & Gore, J. C. (1999). ROC analysis of statistical methods used in functional MRI: individual subjects. NeuroImage, 9: 311329.CrossRefGoogle ScholarPubMed
Smith, S. M. (2012). The future of FMRI connectivity. NeuroImage, 62: 12571266.Google Scholar
Smith, S. M., Jenkinson, M., Beckmann, C., Miller, K., & Woolrich, M. (2007). Meaningful design and contrast estimability in FMRI. NeuroImage, 34: 127136.Google Scholar
Smith, S. M., Jenkinson, M., Woolrich, M. W., Beckmann, C. F., Behrens, T. E., Johansen-Berg, H., … & Matthews, P. M. (2004). Advances in functional and structural MR image analysis and implementation as FSL. NeuroImage, 23: S208S219.Google Scholar
Sporns, O. (2014). Contributions and challenges for network models in cognitive neuroscience. Nature Neuroscience, 17: 652660.Google Scholar
Stephan, K. E., Penny, W. D., Daunizeau, J., Moran, R. J., & Friston, K. J. (2009). Bayesian model selection for group studies. NeuroImage, 46: 10041017.Google Scholar
Sternberg, S. (2001). Separate modifiability, mental modules, and the use of pure and composite measures to reveal them. Acta Psychologica (Amsterdam), 106: 147246.Google Scholar
Summerfield, C., Greene, M., Wager, T., Egner, T., Hirsch, J., & Mangels, J. (2006). Neocortical connectivity during episodic memory formation. PLoS Biol, 4: e128.Google Scholar
Summerfield, C., Trittschuh, E. H., Monti, J. M., Mesulam, M. M., & Egner, T. (2008). Neural repetition suppression reflects fulfilled perceptual expectations. Nature Neuroscience, 11: 10041006.Google Scholar
Sylvester, C. Y., Wager, T. D., Lacey, S. C., Hernandez, L., Nichols, T. E., Smith, E. E., & Jonides, J. (2003). Switching attention and resolving interference: fMRI measures of executive functions. Neuropsychologia, 41: 357370.Google Scholar
Tagliazucchi, E. & Laufs, H. (2014). Decoding wakefulness levels from typical fMRI resting-state data reveals reliable drifts between wakefulness and sleep. Neuron, 82: 695708.Google Scholar
Talairach, J. & Tournoux, P. (1988). Co-Planar Stereotaxic Atlas of the Human Brain. 3-Dimensional Proportional System: An Approach to Cerebral Imaging. Stuttgart and New York: Thieme.Google Scholar
Taylor, J. E. & Worsley, K. J. (2006). Inference for magnitudes and delays of responses in the FIAC data using BRAINSTAT/FMRISTAT. Human Brain Mapping, 27: 434441.Google Scholar
Thompson, P. M., Schwartz, C., Lin, R. T., Khan, A. A., & Toga, A. W. (1996). Three-dimensional statistical analysis of sulcal variability in the human brain. Journal of Neuroscience, 16: 42614274.Google Scholar
Thompson, P. M., Stein, J. L., Medland, S. E., Hibar, D. P., Vasquez, A. A., Renteria, M. E., … & Drevets, W. (2014). The ENIGMA Consortium: large-scale collaborative analyses of neuroimaging and genetic data. Brain Imaging and Behavior, 8: 153182.Google Scholar
Tye, K. M., Prakash, R., Kim, S.-Y., Fenno, L. E., Grosenick, L., Zarabi, H., … & Deisseroth, K. (2011). Amygdala circuitry mediating reversible and bidirectional control of anxiety. Nature, 471: 358362.Google Scholar
van Ast, V., Spicer, J., Smith, E., Schmer-Galunder, S., Liberzon, I., Abelson, J., & Wager, T. (2014). Brain mechanisms of social threat effects on working memory. Cerebral Cortex (September): bhu206.Google Scholar
Van Essen, D. C. & Dierker, D. L. (2007). Surface-based and probabilistic atlases of primate cerebral cortex. Neuron, 56: 209225.Google Scholar
Van Essen, D. C., Drury, H. A., Dickson, J., Harwell, J., Hanlon, D., & Anderson, C. H. (2001). An integrated software suite for surface-based analyses of cerebral cortex. Journal of the American Medical Informatics Association, 8: 443459.Google Scholar
Vazquez, A. L., Cohen, E. R., Gulani, V., Hernandez-Garcia, L., Zheng, Y., Lee, G. R., … & Noll, D. C. (2006). Vascular dynamics and BOLD fMRI: CBF level effects and analysis considerations. NeuroImage, 32: 16421655.Google Scholar
Vazquez, A. L. & Noll, D. C. (1998). Nonlinear aspects of the BOLD response in functional MRI. NeuroImage, 7: 108118.Google Scholar
Vincent, J. L., Patel, G. H., Fox, M. D., Snyder, A. Z., Baker, J. T., Van Essen, D. C., … & Raichle, M. E. (2007). Intrinsic functional architecture in the anaesthetized monkey brain. Nature, 447: 8386.CrossRefGoogle ScholarPubMed
Vogt, B. A., Nimchinsky, E. A., Vogt, L. J., & Hof, P. R. (1995). Human cingulate cortex: surface features, flat maps, and cytoarchitecture. Journal of Comparative Neurology, 359: 490506.Google Scholar
Vul, E., Harris, C., Winkielman, P., & Pashler, H. (2009). Puzzlingly high correlations in fMRI studies of emotion, personality, and social cognition. Perspectives on Psychological Science, 4: 274290.CrossRefGoogle ScholarPubMed
Wager, T. D., Atlas, L. Y., Lindquist, M. A., Roy, M., Woo, C. W., & Kross, E. (2013). An fMRI-based neurologic signature of physical pain. New England Journal of Medicine, 368: 13881397.Google Scholar
Wager, T. D., Jonides, J., & Reading, S. (2004a). Neuroimaging studies of shifting attention: a meta-analysis. NeuroImage, 22: 16791693.Google Scholar
Wager, T. D., Jonides, J., Smith, E. E., & Nichols, T. E. (2005b). Toward a taxonomy of attention shifting: individual differences in fMRI during multiple shift types. Cognitive, Affective, & Behavioral Neuroscience, 5: 127143.Google Scholar
Wager, T. D., Lindquist, M., & Kaplan, L. (2007). Meta-analysis of functional neuroimaging data: Current and future directions. Social Cognitive and Affective Neuroscience, 2: 150158.Google Scholar
Wager, T. D. & Nichols, T. E. (2003). Optimization of experimental design in fMRI: a general framework using a genetic algorithm. NeuroImage, 18: 293309.Google Scholar
Wager, T. D., Reading, S., & Jonides, J. (2004b). Neuroimaging studies of shifting attention: a meta-analysis. NeuroImage, 22: 16791693.Google Scholar
Wager, T. D., Vazquez, A, Hernandez, L, & Noll, D. C. (2005a). Accounting for nonlinear BOLD effects in fMRI: parameter estimates and a model for prediction in rapid event-related studies. NeuroImage, 25: 206218.Google Scholar
Wager, T. D., Waugh, C. E., Lindquist, M., Noll, D. C., Fredrickson, B. L., & Taylor, S. F. (2009). Brain mediators of cardiovascular responses to social threat. Part I: Reciprocal dorsal and ventral sub-regions of the medial prefrontal cortex and heart-rate reactivity. NeuroImage, 47: 821835.Google Scholar
Waugh, C. E., Hamilton, J. P., & Gotlib, I. H. (2010). The neural temporal dynamics of the intensity of emotional experience. NeuroImage, 49: 16991707.CrossRefGoogle ScholarPubMed
Wiech, K., Jbabdi, S., Lin, C. S., Andersson, J., & Tracey, I. (2014). Differential structural and resting state connectivity between insular subdivisions and other pain-related brain regions. Pain, 155: 20472055.Google Scholar
Wilson, J. L. & Jezzard, P. (2003). Utilization of an intra-oral diamagnetic passive shim in functional MRI of the inferior frontal cortex. Magnetic Resonance in Medicine, 50: 10891094.Google Scholar
Winkler, A. M., Ridgway, G. R., Webster, M. A., Smith, S. M., & Nichols, T. E. (2014). Permutation inference for the general linear model. NeuroImage, 92: 381397.Google Scholar
Wise, R. G., Rogers, R., Painter, D., Bantick, S., Ploghaus, A., Williams, P., … & Tracey, I. (2002). Combining fMRI with a pharmacokinetic model to determine which brain areas activated by painful stimulation are specifically modulated by remifentanil. NeuroImage, 16: 9991014.Google Scholar
Woo, C. W., Koban, L., Kross, E., Lindquist, M. A., Banich, M. T., Ruzic, L., … & Wager, T. D. (2014a). Separate neural representations for physical pain and social rejection. Nature Communications, 5: 5380.Google Scholar
Woo, C. W., Krishnan, A., & Wager, T. D. (2014b). Cluster-extent based thresholding in fMRI analyses: pitfalls and recommendations. NeuroImage, 91: 412419.Google Scholar
Woolrich, M. W., Behrens, T. E., Beckmann, C. F., Jenkinson, M., & Smith, S. M. (2004). Multilevel linear modelling for FMRI group analysis using Bayesian inference. NeuroImage, 21: 17321747.Google Scholar
Worsley, K. J. & Friston, K. J. (1995). Analysis of fMRI time-series revisited – again. NeuroImage, 2: 173181.Google Scholar
Worsley, K. J., Taylor, J. E., Tomaiuolo, F., & Lerch, J. (2004). Unified univariate and multivariate random field theory. NeuroImage, 23: S189S195.Google Scholar
Yacubian, J., Sommer, T., Schroeder, K., Glascher, J., Kalisch, R., Leuenberger, B., … & Buchel, C. (2007). Gene–gene interaction associated with neural reward sensitivity. Proceedings of the National Academy of Sciences of the USA, 104: 81258130.Google Scholar
Yarkoni, T., Poldrack, R. A., Nichols, T. E., Van Essen, D. C., & Wager, T. D. (2011). Large-scale automated synthesis of human functional neuroimaging data. Nature Methods, 8: 665670.Google Scholar
Yeo, B. T., Krienen, F. M., Sepulcre, J., Sabuncu, M. R., Lashkari, D., Hollinshead, M., … & Buckner, R. L. (2011). The organization of the human cerebral cortex estimated by intrinsic functional connectivity. Journal of Neurophysiology, 106: 11251165.Google Scholar
Zarahn, E. & Slifstein, M. (2001). A reference effect approach for power analysis in fMRI. NeuroImage, 14: 768779.CrossRefGoogle ScholarPubMed

References

Allison, T., McCarthy, G., Nobre, A., Puce, A., & Belger, A. (1994). Human extrastriate visual cortex and the perception of faces, words, numbers, and colors. Cerebral Cortex, 4: 544554.Google Scholar
Berger, H. (1929). Ueber das Elektrenkephalogramm des Menschen. Archives für Psychiatrie Nervenkrankheiten, 87: 527570.Google Scholar
Brainard, D. H. (1997). The psychophysics toolbox. Spatial Vision, 10: 433436.Google Scholar
Brazdil, M., Roman, R., Falkenstein, M., Daniel, P., Jurak, P., & Rektor, I. (2002). Error processing: evidence from intracerebral ERP recordings. Experimental Brain Research, 146: 460466.Google Scholar
Buzsáki, G., Anastassiou, C. A., & Koch, C. (2012). The origin of extracellular fields and currents: EEG, ECoG, LFP and spikes. Nature Reviews Neuroscience, 13: 407420.Google Scholar
Cheour, M., Leppanen, P. H., & Kraus, N. (2000). Mismatch negativity (MMN) as a tool for investigating auditory discrimination and sensory memory in infants and children. Clinical Neurophysiology, 111: 416.Google Scholar
Cohen, M. X. (2014). Analyzing Neural Time Series Data: Theory and Practice. Cambridge, MA: MIT Press.Google Scholar
Csepe, V. (1995). On the origin and development of the mismatch negativity. Ear and Hearing, 16: 91104.Google Scholar
Dehaene-Lambertz, G. & Baillet, S. (1998). A phonological representation in the infant brain. Neuroreport, 9: 18851888.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, 134: 921.Google Scholar
Eimer, M. & Kiss, M. (2008). Involuntary attentional capture is determined by task set: evidence from event-related brain potentials. Journal of Cognitive Neuroscience, 208: 14231433.Google Scholar
Fischer, C., Luaute, J., Adeleine, P., & Morlet, D. (2004). Predictive value of sensory and cognitive evoked potentials for awakening from coma. Neurology, 63: 669673.Google Scholar
Gehring, W. J., Liu, Y., Orr, J. M., & Carp, J. (2012). The error-related negativity (ERN/Ne). In Luck, S. J. & Kappenman, E. S. (eds.), The Oxford Handbook of Event-Related Potential Components (pp. 231292). Oxford University Press.Google Scholar
Groppe, D. M., Urbach, T. P., & Kutas, M. (2011a). Mass univariate analysis of event-related brain potentials/fields I: a critical tutorial review. Psychophysiology, 48: 17111725.Google Scholar
Groppe, D. M., Urbach, T. P., & Kutas, M. (2011b). Mass univariate analysis of event-related brain potentials/fields II: simulation studies. Psychophysiology, 48: 17261737.Google Scholar
Hopf, J.-M., Luck, S. J., Boelmans, K., Schoenfeld, M. A., Boehler, N., Rieger, J., & Heinze, H.-J. (2006). The neural site of attention matches the spatial scale of perception. Journal of Neuroscience, 26: 35323540.CrossRefGoogle ScholarPubMed
Kappenman, E. S., Farrens, J. L., Luck, S. J., & Hajcak Proudfit, G. (2014). Behavioral and ERP measures of attentional bias to threat in the dot-probe task: poor reliability and lack of correlation with anxiety. Frontiers in Psychology, 5: 1368.Google Scholar
Kappenman, E. S., Kaiser, S. T., Robinson, B. M., Morris, S. E., Hahn, B., Beck, V. M., Leonard, C. J., Gold, J. M., & Luck, S. J. (2012). Response activation impairments in schizophrenia: evidence from the lateralized readiness potential. Psychophysiology, 49: 7384.Google Scholar
Kappenman, E. S. & Luck, S. J. (2010). The effects of electrode impedance on data quality and statistical significance in ERP recordings. Psychophysiology, 47: 888904.Google Scholar
Kappenman, E. S. & Luck, S. J. (2012). ERP components: the ups and downs of brainwave recordings. In Luck, S. J. & Kappenman, E. S. (eds.), The Oxford Handbook of ERP Components (pp. 330). Oxford University Press.Google Scholar
Kappenman, E. S., Luck, S. J., Kring, A. M., Lesh, T. A., Mangun, G. R., Niendam, T., Ragland, J. D., Ranganath, C., Solomon, M., Swaab, T.Y., & Carter, C. S. (2016). Electrophysiological evidence for impaired control of motor output in schizophrenia. Cerebral Cortex, 18911899.Google Scholar
Kappenman, E. S., MacNamara, A., & Hajcak Proudfit, G. (2015). Electrocortical evidence for rapid allocation of attention to threat in the dot-probe task. Social Cognitive and Affective Neuroscience, 10: 577583.Google Scholar
Kayser, J., Tenke, C. E., Bhattacharya, N., Stuart, B. K., Hudson, J., & Bruder, G. E. (2000). Direct comparison of geodesic sensor net (128-channel) and conventional (30-channel) ERPs in tonal and phonetic oddball tasks. Psychophysiology, 37: S17.Google Scholar
Kiesel, A., Miller, J., Jolicoeur, P., & Brisson, B. (2008). Measurement of ERP latency differences: a comparison of single-participant and jackknife-based scoring methods. Psychophysiology, 45: 250274.Google Scholar
Kiss, M., Driver, J., & Eimer, M. (2009). Reward priority of visual target singletons modulates event-related potential signatures of attentional selection. Psychological Science, 20: 245251.Google Scholar
Kutas, M., McCarthy, G., & Donchin, E. (1977). Augmenting mental chronometry: the P300 as a measure of stimulus evaluation time. Science, 197: 792795.Google Scholar
Lopez-Calderon, J. & Luck, S. J. (2014). ERPLAB: an open-source toolbox for the analysis of event-related potentials. Frontiers in Human Neuroscience, 8: 213.Google Scholar
Lorenzo-Lopez, L., Amenedo, E., & Cadaveira, F. (2008). Feature processing during visual search in normal aging: electrophysiological evidence. Neurobiology of Aging, 29: 11011110.Google Scholar
Luck, S. J. (2012). Electrophysiological correlates of the focusing of attention within complex visual scenes: N2pc and related ERP components. In Luck, S. J. & Kappenman, E. S. (eds.), The Oxford Handbook of ERP Components (pp. 329360). Oxford University Press.Google Scholar
Luck, S. J. (2014). An Introduction to the Event-Related Potential Technique, 2nd edn. Cambridge, MA: MIT Press.Google Scholar
Luck, S. J., Fuller, R. L., Braun, E. L., Robinson, B., Summerfelt, A., & Gold, J. M. (2006). The speed of visual attention in schizophrenia: electrophysiological and behavioral evidence. Schizophrenia Research, 85: 174195.Google Scholar
Luck, S. J. & Kappenman, E. S. (eds.) (2012). The Oxford Handbook of Event-Related Potential Components. Oxford University Press.Google Scholar
Luck, S. J., Kappenman, E. S., Fuller, R. L., Robinson, B., Summerfelt, A., & Gold, J. M. (2009). Impaired response selection in schizophrenia: evidence from the P3 wave and the lateralized readiness potential. Psychophysiology, 46: 776786.Google Scholar
Luck, S. J., Mathalon, D. H., O’Donnell, B. F., Spencer, K. M., Javitt, D. C., Ulhaaus, P. F., & Hämäläinen, M. S. (2011). A roadmap for the development and validation of ERP biomarkers in schizophrenia research. Biological Psychiatry, 70: 2834.Google Scholar
Makeig, S. & Onton, J. (2012). ERP features and EEG dynamics: an ICA perspective. In Luck, S. J. & Kappenman, E. S. (eds.), The Oxford Handbook of ERP Components (pp. 5186). Oxford University Press.Google Scholar
Maris, E. & Oostenveld, R. (2007). Nonparametric statistical testing of EEG- and MEG-data. Journal of Neuroscience Methods, 164: 177190.Google Scholar
Näätänen, R. & Kreegipuu, K. (2012). The mismatch negativity (MMN). In Luck, S. J. & Kappenman, E. S. (eds.), The Oxford Handbook of Event-Related Potential Components (pp. 143157). Oxford University Press.Google Scholar
Nunez, P. L. & Srinivasan, R. (2006). Electric Fields of the Brain, 2nd edn. Oxford University Press.Google Scholar
Ochoa, C. J. & Polich, J. (2000). P300 and blink instructions. Clinical Neurophysiology, 111: 9398.Google Scholar
Peirce, J. W. (2007). PsychoPy: psychophysics software in Python. Journal of Neuroscience Methods, 162: 813.Google Scholar
Pelli, D. G. (1997). The VideoToolbox software for visual psychophysics: transforming numbers into movies. Spatial Vision, 10: 437442.Google Scholar
Perez, V. B. & Vogel, E. K. (2012). What ERPs can tell us about working memory. In Luck, S. J. & Kappenman, E. S. (eds.), The Oxford Handbook of Event-Related Potential Components (pp. 361372). Oxford University Press.Google Scholar
Picton, T. W. (2011). Human Auditory Evoked Potentials. San Diego, CA: Plural Publishing.Google Scholar
Polich, J. (2012). Neuropsychology of P300. In Luck, S. J. & Kappenman, E. S. (eds.), The Oxford Handbook of Event-Related Potential Components (pp. 159188). Oxford University Press.Google Scholar
Regan, D. (1989). Human Brain Electrophysiology: Evoked Potentials and Evoked Magnetic Fields in Science and Medicine. New York: Elsevier.Google Scholar
Roach, B. J. & Mathalon, D. H. (2008). Event-related EEG time-frequency analysis: an overview of measures and an analysis of early gamma band phase locking in schizophrenia. Schizophrenia Bulletin, 34: 907926.Google Scholar
Sawaki, R. & Luck, S. J. (2010). Capture versus suppression of attention by salient singletons: electrophysiological evidence for an automatic attend-to-me signal. Attention, Perception, & Psychophysics, 72: 14551470.Google Scholar
Spencer, K. M., Dien, J., & Donchin, E. (2001). Spatiotemporal analysis of the late ERP responses to deviant stimuli. Psychophysiology, 38: 343358.Google Scholar
Tanner, D., Morgan-Short, K., & Luck, S. J. (in press). How inappropriate high-pass filters can produce artifactual effects and incorrect conclusions in ERP studies of language and cognition. Psychophysiology.Google Scholar
Trainor, L., McFadden, M., Hodgson, L., Darragh, L., Barlow, J., Matsos, L., & Sonnadara, R. (2003). Changes in auditory cortex and the development of mismatch negativity between 2 and 6 months of age. International Journal of Psychophysiology, 51: 515.Google Scholar
Woldorff, M. G. (1993). Distortion of ERP averages due to overlap from temporally adjacent ERPs: analysis and correction. Psychophysiology, 30: 98119.Google Scholar
Woldorff, M. G., Hackley, S. A., & Hillyard, S. A. (1991). The effects of channel-selective attention on the mismatch negativity wave elicited by deviant tones. Psychophysiology, 28: 3042.Google Scholar
Yeung, N. (2004). Relating cognitive and affective theories of the error-related negativity. In Ullsperger, M. & Falkenstein, M. (eds.), Errors, Conflicts, and the Brain: Current Opinions on Performance Monitoring (pp. 6370). Leipzig: MPI of Cognitive Neuroscience.Google Scholar

References

Berger, H. (1929). Ueder das Elektroencephalogramm des Menschen. Archives für Psychiatry Nervenkrankheiten, 87: 527570.Google Scholar
Brodbeck, V., Kuhn, A., von Wegner, F., Morzelewski, A. Tagliazucchi, E., Borisov, S., … & Laufs, H. (2012). EEG microstates of wakefulness and NREM sleep. NeuroImage, 62: 21292139.Google Scholar
Brunet, D., Murray, M. M., & Michel, C. M. (2011). Spatio-temporal analysis of multichannel EEG: CARTOOL. Computational Intelligence and Neuroscience, 1: 813870.Google Scholar
Button, K. S., Ioannidis, J. P., Mokrysz, C., Nosek, B. A., Flint, J., Robinson, E. S., & Munafò, M. R. (2013). Empirical evidence for low reproducibility indicates low pre-study odds. Nature Reviews Neuroscience, 14: 877.Google Scholar
Cacioppo, J. T. & Cacioppo, S. (2013). Minimal replicability, generalizability, and scientific advances in psychological science. European Journal of Personality, 27: 121122.Google Scholar
Cacioppo, J. T. & Dorfman, D. D. (1987). Waveform moment analysis in psychophysiological research. Psychological Bulletin, 102: 421438.Google Scholar
Cacioppo, J. T., Tassinary, L. G., & Berntson, G. G. (2000). Handbook of Psychophysiology, 2nd edn. Cambridge University Press.Google Scholar
Cacioppo, S., Balogh, S., & Cacioppo, J. T. (2015). Implicit attention to negative social, in contrast to nonsocial, words in the Stroop task differs between individuals high and low in loneliness: evidence from event-related brain microstates. Cortex, 70: 213233.Google Scholar
Cacioppo, S., Banagee, M., Balogh, S., Cardenas-Iniguez, C., Qualter, P., & Cacioppo, J. T. (2016). Loneliness and implicit attention to social threat: a high performance electrical neuroimaging study. Cognitive Neuroscience, 7: 138159.Google Scholar
Cacioppo, S., Bianchi-Demicheli, F., Bischof, P., Deziegler, E., Michel, C. M., & Landis, T. (2013a). Hemispheric specialization varies with EEG brain resting states and phase of menstrual cycle. PLoS One, 8: e63196.Google Scholar
Cacioppo, S. & Cacioppo, J. T. (2015). Dynamic spatiotemporal brain analyses using high-performance electrical neuroimaging: Part II. A step-by-step tutorial. Journal of Neuroscience Methods, 256: 184197.Google Scholar
Cacioppo, S., Frum, C., Asp, E., Weiss, R. M., Lewis, L. W., & Cacioppo, J. T. (2013b). A quantitative meta-analysis of functional imaging studies of social rejection. Scientific Reports, 3: 2027.Google Scholar
Cacioppo, S., Weiss, R. M., Runesha, H. B., & Cacioppo, J. T. (2014). Dynamic spatiotemporal brain analyses using high-performance electrical neuroimaging: theoretical framework and validation. Journal of Neuroscience Methods, 238: 1134.Google Scholar
Collura, T. F. (1993). History and evolution of electroencephalographic instruments and techniques. Journal of Clinical Neurophysiology, 10: 476504.Google Scholar
Decety, J. & Cacioppo, S. (2012). The speed of morality: a high-density electrical neuroimaging study. Journal of Neurophysiology, 108: 30683072.Google Scholar
Delorme, A., Palmer, J., Onton, J., Oostenveld, R., & Makeig, S. (2012). Independent EEG sources are dipolar. PLoS One, 7: e30135.Google Scholar
Francis, G. (2014). The frequency of excess success for articles in Psychological Science. Psychonomic Bulletin & Review, 21, 11801187.Google Scholar
Gartner, M., Brodbeck, V., Helmut, L., & Schneider, G. (2015). A stochastic model for EEG microstate sequence analysis. NeuroImage, 104: 199208.Google Scholar
Gloor, P. (1969). Hans Berger on the electroencephalogram of man. EEG Clinical Neurophysiology, Suppl. 28: 136.Google Scholar
Khanna, A., Pascual-Leone, A., Michel, C. M., & Farzan, F. (2015). Microstates in resting-state EEG: current status and future directions. Neuroscience & Biobehavioral Reviews, 49: 105113.Google Scholar
Koenig, T., Prichep, L., Lehmann, D., Sosa, P. V., Braeker, E., Kleinlogel, H., … & John, E. R. (2002). Millisecond by millisecond, year by year: normative EEG microstates and developmental stages. NeuroImage, 1: 4148.Google Scholar
Lehmann, D. (1987). Principles of spatial analysis. In Gevins, A. & Remond, A. (eds.), Handbook of Electroencephalography and Clinical Neurophysiology, Vol. 1: Methods of Analysis of Brain Electrical and Magnetic Signals (pp. 309354). Amsterdam: Elsevier.Google Scholar
Lehmann, D. & Skrandies, W. (1980). Reference-free identification of components of checkerboard-evoked multichannel potential fields. Electroencephalography & Clinical Neurophysiology, 48: 609621.Google Scholar
Lehmann, D. & Skrandies, W. (1984). Spatial analysis of evoked potentials in man: a review. Progress in Neurobiology, 23: 227250.Google Scholar
Luck, S. J. (2014). An Introduction to the Event-Related Potential Technique. Cambridge, MA: MIT Press.Google Scholar
Luck, S. J. & Kappenman, E. S. (eds.) (2012). The Oxford Handbook of Event-Related Potential Components. New York: Oxford University Press.Google Scholar
Ortigue, S., Michel, C. M., Murray, M. M., Mohr, C., Carbonnel, S., & Landis, T. (2004). Electrical neuroimaging reveals early generator modulation to emotional words. NeuroImage, 21: 12421251.Google Scholar
Ortigue, S., Sinigaglia, C., Rizzolatti, G., & Grafton, S. T. (2010). Understanding actions of others: the electrodynamics of the left and right hemispheres – a high-density EEG neuroimaging study. PLoS One, 5: 13.Google Scholar
Ortigue, S., Thompson, J. C., Parasuraman, R., & Grafton, S. T. (2009). Spatio-temporal dynamics of human intention understanding in temporo-parietal cortex: a combined EEG/fMRI repetition suppression paradigm. PLoS One, 4: 6962.Google Scholar
Pascual-Marqui, R. D., Michel, C. M., & Lehmann, D. (1995). Segmentation of brain electrical activity into microstates: model estimation and validation. IEEE Transactions on Biomedical Engineering, 42: 658665.Google Scholar
Picton, T. W., Bentin, S., Berg, P., Donchin, E., Hillyard, S. A., Johnson, R. J., … & Taylor, M. J. (2000). Guidelines for using event-related potentials to study cognition: recording standards and publication criteria. Psychophysiology, 37: 127152.Google Scholar
Sarter, M., Bernston, G. G., & Cacioppo, J. T. (1996). Brain imaging and cognitive neuroscience: toward strong inference in attributing function to structure. American Psychologist, 51: 1321.Google Scholar
Tadel, F., Baillet, S., Mosher, J. C., Pantazis, D., & Leahy, R. M. (2011). Brainstorm: a user-friendly application for MEG/EEG analysis. Computational Intelligence and Neuroscience, ID 879716.Google Scholar
Volkmer, T., Tahaghoghi, S. M. M., & Williams, H. E. (2004). Gradual transition detection using average frame similarity. Proceedings of the 2004 Conference on Computer Vision and Pattern Recognition Workshop, vol. 9 (p. 139). Washington, DC: IEEE Computer Society.Google Scholar

References

Abraham, W. C. & Bear, M. F. (1996). Metaplasticity: the plasticity of synaptic plasticity. Trends in Neurosciences, 19: 126130.Google Scholar
Agnew, W. F. & McCreery, D. B. (1987). Considerations for safety in the use of extracranial stimulation for motor evoked potentials. Neurosurgery, 20: 143147.Google Scholar
Agudelo-Toro, A. & Neef, A. (2013). Computationally efficient simulation of electrical activity at cell membranes interacting with self-generated and externally imposed electric fields. Journal of Neural Engineering, 10: 026019.Google Scholar
Albert, D. J. (1966). The effects of polarizing currents on the consolidation of learning. Neuropsychologia, 4: 6577.CrossRefGoogle Scholar
Allen, E. A., Pasley, B. N., Duong, T., & Freeman, R.D. (2007). Transcranial magnetic stimulation elicits coupled neural and hemodynamic consequences. Science, 317: 19181921.Google Scholar
Amassian, V. E., Cracco, R. Q., Maccabee, P. J., Cracco, J. B., Rudell, A., & Eberle, L. (1989). Suppression of visual perception by magnetic coil-stimulation of human occipital cortex. Electroencephalography & Clinical Neurophysiology, 74: 458462.Google Scholar
Amassian, V. E., Maccabee, P. J., Cracco, R. Q., Cracco, J. B., Rudell, A. P., & Eberle, L. (1993). Measurement of information processing delays in human visual cortex with repetitive magnetic coil stimulation. Brain Research, 605: 317321.Google Scholar
Antal, A., Kincses, T. Z., Nitsche, M. A., Bartfai, O., Demmer, I., & Sommer, M. (2002). Pulse configuration-dependent effects of repetitive transcranial magnetic stimulation on visual perception. NeuroReport, 13: 22292233.Google Scholar
Antal, A., Nitsche, M. A., Kincses, T. Z., Lampe, C., & Paulus, W. (2003). No correlation between moving phosphene and motor thresholds: a transcranial magnetic stimulation study. Neuroreport, 15: 297302.Google Scholar
Awiszus, F. (2003). TMS and threshold hunting. Supplements to EEG Clinical Neurophysiology, 56: 1323.Google Scholar
Aydin-Abidin, S., Trippe, J., Funke, K., Eysel, U. T., & Benali, A. (2008). High- and low-frequency repetitive transcranial magnetic stimulation differentially activates c-Fos and zif268 protein expression in the rat brain. Experimental Brain Research, 188: 249261.Google Scholar
Bagati, D., Mittal, S., Praharaj, S. K., Sarcar, M., Kakra, M., & Kurnar, P. (2012). Repetitive transcranial magnetic stimulation safely administered after seizure. Journal of ECT, 28: 6061.Google Scholar
Balslev, D., Braet, W., McAllister, C., & Miall, R. C. (2007). Interindividual variability in optimal current direction for transcranial magnetic stimulation of the motor cortex. Journal of Neuroscience Methods, 162: 309313.Google Scholar
Barker, A. T. (2002). The history and basic principles of magnetic nerve stimulation. In Pascual-Leone, A., Davey, N. J., Rothwell, J., Wassermann, E. M., & Puri, B. K. (eds.), Handbook of Transcranial Magnetic Stimulation (pp. 317). London: Arnold.Google Scholar
Barker, A. T., Garnham, C. W., & Freeston, I. L. (1991). Magnetic nerve stimulation: the effect of waveform on efficiency, determination of neural membrane time constants and the measurement of stimulator output. Electroencephalography & Clinical Neurophysiology Supplement, 43: 227237.Google Scholar
Barker, A. T., Jalinous, R., & Freeston, I. L. (1985). Non-invasive magnetic stimulation of human motor cortex. Lancet, 1: 11061107.Google Scholar
Berardelli, A., Inghilleri, M., Gilio, F., Romeo, S., Pedace, F., Currà, A., & Manfredi, M. (1999). Effects of repetitive cortical stimulation on the silent period evoked by magnetic stimulation. Experimental Brain Research, 125: 8286.Google Scholar
Berardelli, A., Inghilleri, M., Rothwell, J. C., Romeo, S., Currà, A., Gilio, F., … & Manfredi, M. (1998). Facilitation of muscle evoked responses after repetitive cortical stimulation in man. Experimental Brain Research, 122: 7984.Google Scholar
Bestmann, S., Thilo, K. V., Sauner, D., Siebner, H. R., & Rothwell, J.C. (2002). Parietal magnetic stimulation delays visuomotor mental rotation at increased processing demands. NeuroImage, 17: 15121520.CrossRefGoogle ScholarPubMed
Bestmann, S. J., Baudewig, J., & Frahm, J. (2003a). On the synchronization of transcranial magnetic stimulation and functional echo-planar imaging. Journal of Magnetic Resonance Imaging, 17: 309316.Google Scholar
Bestmann, S., Baudewig, J., Siebner, H. R., Rothwell, J. C., & Frahm, J. (2003b). Subthreshold high-frequency TMS of human primary motor cortex modulates interconnected frontal motor areas as detected by interleaved fMRI-TMS. NeuroImage, 20: 16851696.Google Scholar
Bestmann, S., Baudewig, J., Siebner, H. R., Rothwell, J. C., & Frahm, J. (2004). Functional MRI of the immediate impact of transcranial magnetic stimulation on cortical and subcortical motor circuits. European Journal of Neuroscience, 19: 19501962.Google Scholar
Bijsterbosch, J. D., Barker, A. T., Lee, K.-H., & Woodruff, P. W. R. (2012). Where does transcranial magnetic stimulation (TMS) stimulate? Modelling of induced field maps for some common cortical and cerebellar targets. Medical Biological Engineering and Computing, 50: 671681.Google Scholar
Bikson, M., Bulow, P., Stiller, J. W., Datta, A., Battaglia, F., Karnup, S. V., & Postolache, T. T. (2008). Transcranial direct current stimulation for major depression: a general system for quantifying transcranial electrotherapy dosage. Current Treatment Options in Neurology, 10: 377385.Google Scholar
Bliss, T. V., Collingridge, G. L., & Morris, R. G. (2003) Introduction: longterm potentiation and structure of the issue. Philosophical Transactions of the Royal Society London B: Biological Sciences, 358: 607611.Google Scholar
Bliss, T. V. & Lomo, T. (1973). Long-lasting potentiation of synaptic transmission in the dentate area of the anaesthetized rabbit following stimulation of the perforant path. Journal of Physiology, 232: 331356.Google Scholar
Bohning, D. E., Shastri, A., Nahas, Z., Lorberbaum, J. P., Andersen, S. W., Dannels, W. R., … & George, M. S. (1998). Echopolar BOLD fMRI of brain activation induced by concurrent transcranial magnetic stimulation. Investigative Radiology, 33: 336340.Google Scholar
Bonato, C., Miniussi, C., & Rossini, P. M. (2006).Transcranial magnetic stimulation and cortical evoked potentials: a TMS/EEG co-registration study. Clinical Neurophysiology, 117: 16991707.Google Scholar
Boroojerdi, B., Battaglia, F., Muellbacher, W., & Cohen, W. G. (2001a). Mechanisms influencing stimulus-response properties of the human corticospinal system. Clinical Neurophysiology, 112: 931937.Google Scholar
Boroojerdi, B., Phipps, M., Kopylev, L., Wharton, C. M., Cohen, L. G., & Grafman, J. (2001b). Enhancing analogic reasoning with rTMS over the left prefrontal cortex. Neurology, 56: 526528.Google Scholar
Boroojerdi, B., Prager, A., Muellbacher, W., & Cohen, L. G. (2000). Reduction of human visual cortex excitability using 1 Hz transcranial magnetic stimulation. Neurology, 54: 15291531.Google Scholar
Brasil-Neto, J. P., McShane, L. M., Fuhr, P., Hallett, M., & Cohen, L. G. (1992). Topographic mapping of the human cortex with magnetic stimulation: factors affecting accuracy and reproducibility. Electroencephalography & Clinical Neurophysiology, 85: 916.Google Scholar
Brem, A.-K., Fried, P. J., Horvath, J. C., Robertson, E. M., & Pascual-Leone, A. (2014). Is neuroenhancement by noninvasive brain stimulation a zero-sum proposition? NeuroImage, 85: 10581068.Google Scholar
Brett, M., Johnsrude, I. S., & Owen, A. M. (2002). The problem of functional localization in the human brain. Nature Reviews Neuroscience, 3: 243249.Google Scholar
Bungert, A., Chambers, C. D., Phillips, M., & Evans, J. (2006). Reducing image artefacts in concurrent TMS/fMRI by passive shimming. NeuroImage, 59: 21672174.Google Scholar
Buzsaki, G. (2006). Rhythms of the Brain. Oxford University Press.Google Scholar
Canolty, R. T., Edwards, E., Dalal, S. S., Soltani, M., Nagarajan, S. S., Kirsch, H. E., … & Knight, R. T. (2006). High gamma power is phase-locked to theta oscillations in human neocortex. Science, 313: 16261628.Google Scholar
Cantarero, G. & Celnik, P. (2015). Applications of TMS to study brain connectivity. In Reti, I. M. (ed.), Brain Stimulation: Methodologies and Interventions (pp. 191212). Hoboken, NJ: Wiley-Blackwell.Google Scholar
Caparelli, E. C., Backus, W., Telang, F., Wang, G.-J., Maloney, T., Goldstein, R. Z., & Henn., F. (2012). Is 1 Hz rTMS always inhibitory in healthy individuals? The Open Neuroimaging Journal, 6: 6974.Google Scholar
Cardenas-Morales, L., Nowak, D. A., Kammer, T., Wolf, R. C., & Schonfeldt-Lecuona, C. (2010). Mechanisms and applications of theta-burst rTMS on the human motor cortex. Brain Topography, 22: 294306.Google Scholar
Cattaneo, Z., Rota, F., Vecchi, T., & Silvanto, J. (2008). Using state-dependency of transcranial magnetic stimulation (TMS) to investigate letter selectivity in the left posterior parietal cortex: a comparison of TMS-priming and TMS-adaptation paradigms. European Journal of Neuroscience, 28: 19241929.Google Scholar
Cattaneo, Z., Rota, F., Walsh, V., Vecchi, T., & Silvanto, J. (2009). TMS-adaptation reveals abstract letter selectivity in the left posterior parietal cortex. Cerebral Cortex, 19: 23212325.Google Scholar
Chanes, L., Quentin, R., Tallon-Baudry, C., & Valero-Cabré, A. (2013). Causal frequency-specific contributions of frontal spatiotemporal patterns induced by noninvasive neurostimulation to human visual performance. Journal of Neuroscience, 33: 50005005.Google Scholar
Cheeran, B., Talelli, P., Mori, F., Koch, G., Suppa, A., Edwards, M., … & Rothwell, J. C. (2008). A common polymorphism in the brain-derived neurotrophic factor gene (BDNF) modulates human cortical plasticity and the response to rTMS. Journal of Physiology, 586: 57175725.Google Scholar
Chen, A. C., Oathes, D. J., Chang, C., Bradley, T., Zhou, Z. W., Williams, L. M., … & Etkin, A. (2013). Causal interactions between fronto-parietal central executive and default-mode networks in humans. Proceedings of the National Academy of Sciences of the USA, 110: 1994419949.Google Scholar
Chen, R. (2004). Interactions between inhibitory and excitatory circuits in the human motor cortex. Experimental Brain Research, 154: 110.Google Scholar
Chen, R., Classen, J., Gerloff, C., Celnik, P., Wassermann, E. M., Hallett, M., & Cohen, L. G. (1997). Depression of motor cortex excitability by low-frequency transcranial magnetic stimulation. Neurology, 48: 13981403.Google Scholar
Chervyakov, A. V., Piradov, M. A., Chernikova, L. A., Nazarova, M. A., Gnezditsky, V. V., Savitskaya, N. G., & Fedin, P. A. (2013). Capability of navigated repeated transcranial magnetic stimulation in stroke rehabilitation (randomized blind sham-controlled study). Journal of the Neurological Sciences, 333: 246247.Google Scholar
Chiramberro, M., Lindberg, N., Isometsä, E., Kähkönen, S., & Appelberg, B. (2013). Repetitive transcranial magnetic stimulation induced seizures in an adolescent patient with major depression: a case report. Brain Stimulation, 6: 830831.Google Scholar
Cohen, D. & Cuffin, B. N. (1991). Developing a more focal magnetic stimulator. Part I: Some basic principles. Journal of Clinical Neurophysiology, 8: 102111.Google Scholar
Cohen, D. A., Freitas, C., Tormos, J. M., Oberman, L., Eldaief, M., & Pascual-Leone, A. (2010). Enhancing plasticity through repeated rTMS sessions: the benefits of a night of sleep. Clinical Neurophysiology, 121: 21592164.Google Scholar
Cohen Kadosh, R., Cohen Kadosh, K., Schuhmann, T., Kaas, A., Goebel, R., Henik, A., & Sack, A. T. (2007). Virtual dyscalculia induced by parietal-lobe TMS impairs automatic magnitude processing. Current Biology, 17: 689693.Google Scholar
Cooper, A. C. G., Humphreys, G. W., Hulleman, J., Praamstra, P., & Georgeson, M. (2004). Transcranial magnetic stimulation to right parietal cortex modifies the attentional blink. Experimental Brain Research, 155: 2429.Google Scholar
Counter, S. A., Borg, E., Lofqvist, L., & Brismar, T. (1990). Hearing loss from the acoustic artifact of the coil used in extracranial magnetic stimulation. Neurology, 40: 11591162.Google Scholar
Crowther, L. J., Porzig, K., Hadimani, R. L., Brauer, H., & Jile, D. C. (2012). Calculation of Lorentz forces on coils for transcranial magnetic stimulation during magnetic resonance imaging. IEEE Transactions on Magnetics, 48: 40584061.Google Scholar
Daskalakis, Z. J., Moller, B., Christensen, B. K., Fitzgerald, P. B., Gunraj, C., & Chen, R. (2006). The effects of repetitive transcranial magnetic stimulation on cortical inhibition in healthy human subjects. Experimental Brain Research, 174: 403412.Google Scholar
Datta, A., Bansal, V., Diaz, J., Patel, J., Reato, D., & Bikson, M. (2009). Gyri-precise head model of transcranial direct current stimulation: improved spatial focality using a ring electrode versus conventional rectangular pad. Brain Stimulation, 2: 201207.Google Scholar
Davey, K. R. & Riehl, M. E. (2006). Suppressing the surface field during transcranial magnetic stimulation. IEEE Transactions on Biomedical Engineering, 53: 190194.CrossRefGoogle ScholarPubMed
de Graaf, T. A., Jacobs, C., Roebroeck, A., & Sack, A. T. (2009). FMRI effective connectivity and TMS chronometry: complementary accounts of causality in the visuospatial judgment network. PLoS One, 4: e8307.Google Scholar
Deblieck, C., Thompson, B., Iacoboni, M., & Wu, A. D. (2008). Correlation between motor and phosphene thresholds: a transcranial magnetic stimulation study. Human Brain Mapping, 29: 662670.Google Scholar
Deng, Z.-D., Lisanby, S. H., & Peterchev, A. V. (2013). Electric field depth–focality tradeoff in transcranial magnetic stimulation: simulation comparison of 50 coil designs. Brain Stimulation, 6: 113.Google Scholar
Deng, Z.-D., Lisanby, S. H., & Peterchev, A. V. (2014). Coil design considerations for deep transcranial magnetic stimulation. Clinical Neurophysiology, 125: 12021212.Google Scholar
Deng, Z.-D. & Peterchev, A. V. (2011). Transcranial magnetic stimulation coil with electronically switchable active and sham modes. Conference Proceedings of the IEEE Engineering Medicine Biology Society.Google Scholar
Denslow, S., Lomarev, M., George, M. S., & Bohning, D. E. (2005). Cortical and subcortical brain effects of transcranial magnetic stimulation (TMS)-induced movement: an interleaved TMS/functional magnetic resonance imaging study. Biological Psychiatry, 57: 752760.CrossRefGoogle ScholarPubMed
Devanne, H., Lavoie, B. A., & Capaday, C. (1997). Input–output properties and gain changes in the human corticospinal pathway. Experimental Brain Research, 114: 329338.Google Scholar
Di Lazzaro, V., Oliviero, A., Profice, P., Pennisi, M. A., Pilato, F., Zito, G., … & Tonali, P. A. (2003). Ketamine increases human motor cortex excitability to transcranial magnetic stimulation. Journal of Physiology, 547: 485496.Google Scholar
Di Lazzaro, V., Pilato, F., Saturno, E., Oliviero, A., Dileone, M., Mazzone, P., … & Rothwell, J. (2005). Theta-burst repetitive transcranial magnetic stimulation suppresses specific excitatory circuits in the human motor cortex. Journal of Physiology, 565: 945950.Google Scholar
Di Lazzaro, V., Profice, P., Ranieri, F., Capone, F., Dileone, M., Oliviero, A., & Pilato, F. (2012). I-wave origin and modulation. Brain Stimulation, 5: 512525.Google Scholar
Dmochowski, J. P., Datta, A., Bikson, M., Su, Y., & Parra, L. C. (2011). Optimized multi-electrode stimulation increases focality and intensity at target. Journal of Neural Engineering, 8: 046011.Google Scholar
Edwardson, M., Fetz, E. E., & Avery, D. H. (2011). Seizure produced by 20 Hz transcranial magnetic stimulation during isometric muscle contraction in a healthy subject. Clinical Neurophysiology, 122: 23242327.Google Scholar
Ellison, A., Battelli, L., Cowey, A., & Walsh, V. (2003). The effect of expectation on facilitation of color/form conjunction tasks by TMS over area V5. Neuropsychologia, 41: 17941801.Google Scholar
Emara, T. H., Moustafa, R. R., Elnahas, N. M., Elganzoury, A. M., Abdo, T. A., Mohamed, S. A., & Eletribi, M. A. (2010). Repetitive transcranial magnetic stimulation at 1 Hz and 5 Hz produces sustained improvement in motor function and disability after ischaemic stroke. European Journal of Neurology, 17: 12031209.Google Scholar
Esser, S. K., Hill, S. L., & Tononi, G. (2005). Modeling the effects of transcranial magnetic stimulation on cortical units. Journal of Neurophysiology, 94: 622639.Google Scholar
Esser, S. K., Huber, R., Massimini, M., Peterson, M. J., Ferrarelli, F., & Tononi, G. (2006). A direct demonstration of cortical LTP in humans: a combined TMS/EEG study. Brain Research Bulletin, 69: 8694.Google Scholar
Feredoes, E., Tononi, G., & Postle, B. R. (2007). The neural bases of the short-term storage of verbal information are anatomically variable across individuals. Journal of Neuroscience, 27: 1100311008.Google Scholar
Ferreri, F. & Rossini, P. M. (2013). TMS and TMS–EEG techniques in the study of the excitability, connectivity, and plasticity of the human motor cortex. Reviews in the Neurosciences, 24: 431442.Google Scholar
Fitzgerald, P. B., Fountain, S., & Daskalakis, Z. J. (2006). A comprehensive review of the effects of rTMS on motor cortical excitability and inhibition. Clinical Neurophysiology, 117: 25842596.Google Scholar
Fox, J. J. & Schroeder, C. E. (2005). The case for feedforward multisensory convergence during early cortical processing. Neuroreport, 16: 419423.Google Scholar
Fox, M. D., Liu, H., & Pascual-Leone, A. (2012). Identification of reproducible individualized targets for treatment of depression with TMS based on intrinsic connectivity. NeuroImage, 66: 151160.Google Scholar
Fox, P., Ingham, R., George, M. S., Mayberg, H., Ingham, J., Roby, J., … & Jerabek, P. (1997). Imaging human intra-cerebral connectivity by PET during TMS. Neuroreport, 8: 27872791.Google Scholar
Freunberger, R., Werkle-Bergner, M., Griesmayr, B., Lindenberger, U., & Klimesch, W. (2011). Brain oscillatory correlates of working memory constraints. Brain Research, 1375: 93102.Google Scholar
Fridlund, A. J. & Cacioppo, J. T. (1986). Guidelines for human electromyographic research. Psychophysiology, 23: 567589.Google Scholar
Fuggetta, G., Pavone, E. F., Fiaschi, A., & Manganotti, P. (2008). Acute modulation of cortical oscillatory activities during short trains of high-frequency repetitive transcranial magnetic stimulation of the human motor cortex: a combined EEG and TMS study. Human Brain Mapping, 29: 113.Google Scholar
Gangitano, M., Valero-Cabré, A., Tormos, J. M., Mottaghy, F. M., Romero, J. R., & Pascual-Leone, A. (2002). Modulation of input–output curves by low and high frequency repetitive transcranial magnetic stimulation of the motor cortex. Clinical Neurophysiology, 113: 12491257.Google Scholar
George, M. S., Short, E. B., Kerns, S. E., Li, X., Hanlon, C., Pelic, C., … & Fox, J. (2015). Therapeutic applications of rTMS for psychiatric and neurological conditions. In Reti, I. M. (ed.), Brain Stimulation: Methodologies and Interventions (pp. 213232). Hoboken, NJ: Wiley-Blackwell.Google Scholar
Goetz, S. M., Luber, B., Lisanby, S. H., Murphy, D. L., Kozyrkov, I. C., Grill, W., & Peterchev, A. V. (2016). Enhancement of rTMS neuromodulatory effects with novel waveforms demonstrated via controllable pulse parameter TMS. Brain Stimulation 9: 3947.Google Scholar
Goetz, S. M., Luber, B., Lisanby, S. H., & Peterchev, A. V. (2014). A novel model incorporating two variability sources for describing motor evoked potentials. Brain Stimulation, 7: 541552.Google Scholar
Goetz, S. M., Pfaeffl, M., Huber, J., Singer, M., Marquardt, R., & Weyh, T. (2012). Circuit topology and control principle for a first magnetic stimulator with fully controllable waveform. Conference Proceedings of IEEE Engineering Medical Biology Society.Google Scholar
Goetz, S. M., Whiting, P. A., & Peterchev, A. V. (2011). Threshold estimation with transcranial magnetic stimulation: algorithm comparison. Clinical Neurophysiology, 122: S197.Google Scholar
Gómez, L., Morales, L., Trápaga, O., & Morales, H. (2011). Seizure induced by sub-threshold 10-Hz rTMS in a patient with multiple risk factors. Clinical Neurophysiology, 122: 10571058.Google Scholar
Grefkes, C., Nowak, D. A., Wang, L. E., Dafotakis, M., Eickhoff, S. B., & Fink, G. R. (2010). Modulating cortical connectivity in stroke patients by rTMS assessed with fMRI and dynamic causal modeling. NeuroImage, 50: 233242.Google Scholar
Groppa, S., Oliviero, A., Eisen, A., Quartarone, A., Cohen, L. G., Mall, V., … & Siebner, H. R. (2012). A practical guide to diagnostic transcranial magnetic stimulation: report of an IFCN committee. Clinical Neurophysiology, 123: 858882.Google Scholar
Grosbras, M. H. & Paus, T. (2002). Transcranial magnetic stimulation of the human frontal eye field: effects on visual perception and attention. Journal of Cognitive Neuroscience, 14: 11091120.Google Scholar
Grosbras, M. H. & Paus, T. (2003). Transcranial magnetic stimulation of the human frontal eye field facilitates visual awareness. European Journal of Neuroscience, 18: 31213126.Google Scholar
Hamada, M., Hanajima, R., Terao, Y., Arai, N., Furubayashi, T., Inomata-Terada, S., … & Ugawa, Y. (2007). Quadro-pulse stimulation is more effective than paired-pulse stimulation for plasticity induction of the human motor cortex. Clinical Neurophysiology, 118: 26722682.Google Scholar
Hamada, M., Strigaro, G., Murase, N., Sadnicka, A., Galea, J. M., Edwards, M. J., & Rothwell, J. C. (2012). Cerebellar modulation of human associative plasticity. Journal of Physiology, 590: 23652374.Google Scholar
Hamada, M., Terao, Y., Hanajima, R., Shirota, Y., Nakatani-Enomoto, S., Furubayashi, T., … & Ugawa, Y. (2008). Bidirectional long-term motor cortical plasticity and metaplasticity induced by quadripulse transcranial magnetic stimulation. Journal of Physiology, 586: 39273947.Google Scholar
Hamada, M. & Ugawa, Y. (2010). Quadripulse stimulation:a new patterned rTMS. Restorative Neurology and Neuroscience, 28: 419424.Google Scholar
Hamidi, M., Johnson, J. S., Feredoes, E., & Postle, B. R. (2011). Does high-frequency repetitive transcranial magnetic stimulation produce residual and/or cumulative effects within an experimental session? Brain Topography, 23: 355367.Google Scholar
Hamidi, M., Slagter, H. A., Tononi, G., & Postle, B. R. (2009). Repetitive transcranial magnetic stimulation affects behavior by biasing endogenous cortical oscillations. Frontiers in Integrative Neuroscience, 3. doi: 10.3389/neuro.07.014.2009.Google Scholar
Hannula, H., Neuvonen, T., Savolainen, P., Hiltunen, J., Ma, Y.-Y., Antila, H., … & Pertovaara, A. (2010). Increasing top-down suppression from prefrontal cortex facilitates tactile working memory. NeuroImage, 49: 10911098.Google Scholar
Harel, E. V., Zangen, A., Roth, Y., Reti, I., Braw, Y., & Levkovitz, Y. (2011). H-coil repetitive transcranial magnetic stimulation for the treatment of bipolar depression: an add-on, safety and feasibility study. World Journal of Biological Psychiatry, 12: 119126.Google Scholar
Harris, J. A., Clifford, C. W. G., & Miniussi, C. (2008). The functional effect of transcranial magnetic stimulation: signal suppression or neural noise generation? Journal of Cognitive Neuroscience, 20: 734740.Google Scholar
Haug, B. A., Schönle, P. W., Knobloch, C., & Köhne, M. (1992). Silent period measurement revives as a valuable diagnostic tool with transcranial magnetic stimulation. Electroencephalography & Clinical Neurophysiology, 85: 158160.Google Scholar
Heller, L. & van Hulsteyn, D. B. (1992). Brain stimulation using electromagnetic sources: theoretical aspects. Biophysical Journal, 63: 129138.Google Scholar
Herwig, U., Abler, B., Schonfeldt-Lecuona, C., Wunderlich, A., Grothe, J., Spitzer, M., & Walter, H. (2003a). Verbal storage in a premotor-parietal network: evidence from fMRI-guided magnetic stimulation. NeuroImage, 20: 10321041.Google Scholar
Herwig, U., Satrapi, P., & Schonfeldt-Lecuona, C. (2003b). Using the International 10–20 EEG system for positioning of transcranial magnetic stimulation. Brain Topography, 16: 9599.Google Scholar
Hoeft, F., Wu, D.-A., Hernandez, A., Glover, G. H., & Shimojo, S. (2008). Electronically switchable sham transcranial magnetic stimulation (TMS) system. PLoS One, 3: e1923.Google Scholar
Hoogendam, J. M., Ramakers, G. M. J., & Di Lazzaro, V. (2010). Physiology of repetitive transcranial magnetic stimulation of the human brain. Brain Stimulation, 3: 95118.Google Scholar
Horvath, J. C., Forte, J. D., & Carter, O. (2015). Evidence that transcranial direct current stimulation (tDCS) generates little-to-no reliable neurophysiologic effect beyond MEP amplitude modulation in healthy human subjects: a systematic review. Neuropsychologia, 66: 213236.Google Scholar
Houdayer, E., Degardin, A., Cassim, F., Bocquillon, P., Derambure, P., & Devanne, H. (2008). The effects of low- and high-frequency repetitive TMS on the input/output properties of the human corticospinal pathway. Experimental Brain Research, 187: 207217.Google Scholar
Hu, S. H., Wang, S. S., Zhang, M. M., Wang, J.-W., Hu, J.-B., & Huang, M.-L. (2011). Repetitive transcranial magnetic stimulation-induced seizure of a patient with adolescent-onset depression: a case report and literature review. Journal of International Medical Research, 39: 20392044.Google Scholar
Huang, Y. Z., Chen, R. S., Rothwell, J. C., & Wen, H.-Y. (2007). The after-effect of human theta burst stimulation is NMDA receptor dependent. Clinical Neurophysiology, 118: 1028.Google Scholar
Huang, Y. Z., Edwards, M. J., Rounis, E., Bhalia, K. P., & Rothwell, J. C. (2005). Theta burst stimulation of the human motor cortex. Neuron, 45: 201206.Google Scholar
Huang, Y.-Z., Sommer, M., Thickbroom, G., Hamada, M., Pascual-Leone, A., Paulus, W., … & Ugawa, Y. (2009). Consensus: new methodologies for brain stimulation. Brain Stimulation, 2: 213.Google Scholar
Ilmoniemi, R. J. & Kičić, D. (2010) Methodology for combined TMS and EEG. Brain Topography, 22: 233248.Google Scholar
Ilmoniemi, R. J., Virtanen, J., Ruohonen, J., Karhu, J., Aronen, H. J., Naatanen, R., & Katila, T. (1997). Neuronal responses to magnetic stimulation reveal cortical reactivity and connectivity. Neuroreport, 8: 35373540.Google Scholar
Iriki, A., Pavlides, C., Keller, A., & Asanuma, H. (1989). Long-term potentiation in the motor cortex. Science, 245: 13851387.Google Scholar
Iyer, M. B., Schleper, N., & Wassermann, E. M. (2003). Priming stimulation enhances the depressant effect of low-frequency repetitive transcranial magnetic stimulation. Journal of Neuroscience, 23: 1086710872.Google Scholar
Jalinous, R. (2002). Principles of magnetic stimulator design. In Pascual-Leone, A., Davey, N. J., Rothwell, J., Wassermann, E. M., and Puri, B. K. (eds.), Handbook of Transcranial Magnetic Stimulation (pp. 3038). London: Arnold.Google Scholar
Janicak, P. G., O’Reardon, J. P., Sampson, S. M., Husain, M. M., Lisanby, S. H., Rado, J. T., … & Demitrack, M. A. (2008). Transcranial magnetic stimulation in the treatment of major depressive disorder: a comprehensive summary of safety experience from acute exposure, extended exposure, and during reintroduction treatment. Journal of Clinical Psychiatry, 69: 222232.Google Scholar
Jennum, P., Winkel, H., & Fuglsang-Frederiksen, A. (1995). Repetitive magnetic stimulation and motor evoked potentials. Electroencephalography & Clinical Neurophysiology, 97: 96101.Google Scholar
Ji, R. R., Schlaepfer, T. E., Aizenman, C. D., Epstein, C. M., Qiu, D., Huang, J. C., & Rupp, F. (1998). Repetitive transcranial magnetic stimulation activates specific regions in rat brain. Proceedings of the National Academy of Sciences of the USA, 95: 1563515640Google Scholar
Juan, C. H. & Walsh, V. (2003). Feedback to V1: a reverse hierarchy in vision. Experimental Brain Research, 150: 259263.Google Scholar
Jung, P. & Ziemann, U. (2009). Homeostatic and nonhomeostatic modulation of learning in human motor cortex. Journal of Neuroscience, 29: 55975604.Google Scholar
Kähkönen, S., Komssi, S., Wilenius, J., & Ilmoniemi, R. J. (2005). Prefrontal transcranial magnetic stimulation produces intensity-dependent EEG responses in humans. NeuroImage, 24: 955960.Google Scholar
Kähkönen, S., Wilenius, J., Komssi, S., & Ilmoniemi, R. J. (2004). Distinct differences in cortical reactivity of motor and prefrontal cortices to magnetic stimulation. Clinical Neurophysiology, 115: 583588.Google Scholar
Kaminski, J. A., Korb, F. M., Viliringer, A., & Ott, D. V. M. (2011). Transcranial magnetic stimulation intensities in cognitive paradigms. PloS One, 6: e24836.Google Scholar
Kamitani, Y. & Schimojo, S. (1999). Manifestation of scotomas created by transcranial magnetic stimulation of human visual cortex. Nature Neuroscience, 2: 767771.Google Scholar
Kammer, T., Beck, S., Thielscher, A., Laubis-Herrmann, U., & Topka, H. (2001). Motor thresholds in humans: a transcranial magnetic stimulation study comparing different pulse waveforms, current directions and stimulator types. Clinical Neurophysiology, 112: 250258.Google Scholar
Kammer, T. & Nusseck, H. G. (1998). Are recognition deficits following occipital lobe TMS explained by raised detection thresholds? Neuropsychologia, 36: 11611166.Google Scholar
Kammer, T., Puls, K., Erb, M., & Grodd, W. (2005). Transcranial magnetic stimulation in the visual system. II: Characterization of induced phosphenes and scotomas. Experimental Brain Research, 160: 129140.Google Scholar
Kessler, S. K., Turkeltaub, P. E., Benson, J. G., & Hamilton, R. H. (2012). Differences in the experience of active and sham transcranial direct current stimulation. Brain Stimulation, 5: 155162.Google Scholar
Kimiskidis, V. K., Papagiannopoulos, S., Sotirakoglou, K., Kazis, D. A., Kazis, A., & Mills, K. R. (2005). Silent period to transcranial magnetic stimulation: construction and properties of stimulus–response curves in healthy volunteers. Experimental Brain Research, 163: 2131.Google Scholar
Kleim, J., Chan, S., Pringle, E., Schallert, K., Procaccio, V., Jimenez, R., & Cramer, S. C. (2006). BDNF val66met polymorphism is associated with modified experience-dependent plasticity in human motor cortex. Nature Neuroscience, 9: 735737.Google Scholar
Klimesch, W., Sauseng, P., & Gerloff, C. (2003). Enhancing cognitive performance with repetitive transcranial magnetic stimulation at human individual alpha frequency. European Journal of Neuroscience, 17: 11291133.Google Scholar
Kohler, S., Paus, T., Buckner, R. L., & Milner, B. (2006). Effect of left inferior prefrontal stimulation on episodic memory formation: a two-stage fMRI-rTMS study. Journal of Cognitive Neuroscience, 16: 178188.Google Scholar
Lakatos, P., Shah, A. S., Knuth, K. H., Ulbert, I., Karmos, G., & Shroeder, C. E. (2005). An oscillatory hierarchy controlling neuronal excitability and stimulus processing in the auditory cortex. Journal of Neurophysiology, 94: 19041911.Google Scholar
Lang, N., Rothkegel, H., Reiber, H., Hasan, A., Sueske, E., Tergau, F., … & Paulus, W. (2011). Circadian modulation of GABA-mediated cortical inhibition. Cerebral Cortex, 21: 22992306.Google Scholar
Lang, N., Siebner, H. R., Ernst, D., Nitsche, M. A., Paulus, W., Lemon, R. N., & Rothwell, J. C. (2004). Preconditioning with transcranial direct current stimulation sensitizes the motor cortex to rapid-rate transcranial magnetic stimulation and controls the direction of after-effects. Biological Psychiatry, 56: 634638.Google Scholar
Lang, N., Speck, S., Harms, J., Rothkegel, H., Paulus, W., & Sommer, M. (2008). Dopaminergic potentiation of rTMS induced motor cortex inhibition. Biological Psychiatry, 3: 231233.Google Scholar
Larson, J., Wong, D., & Lynch, G. (1986). Patterned stimulation at the theta frequency is optimal for the induction of hippocampal long-term potentiation. Brain Research, 368: 347350.Google Scholar
Levkovitz, Y., Roth, Y., Harel, E. V., Braw, Y., Sheer, A., & Zangen, A. (2007). A randomized controlled feasibility and safety study of deep transcranial magnetic stimulation. Clinical Neurophysiology, 118: 27302744.Google Scholar
Li, L. M., Uehara, K., & Hanakawa, T. (2015). The contribution of interindividual factors to variability of response in transcranial direct current stimulation studies. Frontiers in Cellular Neuroscience, 9: article 181.Google Scholar
Lisanby, S. H., Gutman, D., Luber, B., Schroeder, C., & Sackeim, H. A. (2001). Sham TMS: intracerebral measurements of the induced electrical field and the induction of motor-evoked potentials. Biological Psychiatry, 49: 460463.Google Scholar
Lisman, J. E. & Jensen, O. (2013). The theta-gamma neural code. Neuron, 77: 10021016.Google Scholar
Liston, C., Chen, A. C., Zebley, B. D., Drysdale, A. T., Gordon, R., Leuchter, B., … & Dubin, M. J. (2014). Default mode network mechanisms of transcranial magnetic stimulation in depression. Biological Psychiatry, 76: 517526.Google Scholar
Lolas, F. (1977). Low-level electric currents and brain indicators of behavioral activation. Arquivos de Neuro-Psiquiatria, 35: 325328.Google Scholar
Loo, C., Sachdev, P., Elsayed, H., McDarmont, B., Mitchell, P., Wilkinson, M., … & Gandevia, S. (2001). Effects of a 2- to 4-week course of repetitive transcranial magnetic stimulation (rTMS) on neuropsychologic functioning, electroencephalogram, and auditory threshold in depressed patients. Biological Psychiatry, 49: 615623.Google Scholar
Lorenzano, C., Gilio, F., Inghilleri, M., & Berardelli, A. (2002). Spread of electrical activity at cortical level after repetitive magnetic stimulation in normal subjects. Experimental Brain Research, 147: 186192.Google Scholar
Luber, B. (2014). Neuroenhancement by noninvasive brain stimulation is not a net zero-sum proposition. Frontiers in Systems Neuroscience, 8: article 129.Google Scholar
Luber, B., Balsam, P., Nguyen, T., Gross, M., & Lisanby, S. H. (2007a). Classical conditioned learning using transcranial magnetic stimulation. Experimental Brain Research, 183: 361369.Google Scholar
Luber, B., Kinnunen, L. H., Rakitin, B. C., Ellsasser, R., Stern, Y., & Lisanby, S. H. (2007b). Facilitation of performance in a working memory task with rTMS stimulation of the precuneus: frequency and time-dependent effects. Brain Research, 1128: 120129.Google Scholar
Luber, B. & Lisanby, S. H. (2014). Enhancement of human cognitive performance using transcranial magnetic stimulation (TMS). NeuroImage, 85: 961970.Google Scholar
Luber, B., Stanford, A. D., Bulow, P., Nguyen, T., Rakitin, B. C., Habeck, C., … & Lisanby, S. H. (2008). Remediation of sleep-deprivation induced visual working memory impairment with fMRI-guided transcranial magnetic stimulation. Cerebral Cortex, 18: 20772085.Google Scholar
Luber, B., Steffener, J., Tucker, A., Habeck, C., Peterchev, A. V., Deng, Z.-D., … & Lisanby, S. H. (2013). Extended remediation of sleep deprivation-induced working memory deficits using fMRI-guided repetitive transcranial magnetic stimulation. Sleep, 36: 857871.Google Scholar
Maccabee, P. J., Amassian, V. E., Eberle, L. P., & Cracco, R. Q. (1993). Magnetic coil stimulation of straight and bent amphibian and mammalian peripheral nerve in vitro: locus of excitation. Journal of Physiology in London, 460: 210219.Google Scholar
Maccabee, P. J., Nagarajan, S. S., Amassian, V. E., Durand, D. M., Szabo, A. Z., Ahad, A. B., … & Eberle, L. P. (1998). Influence of pulse sequence, polarity and amplitude on magnetic stimulation of human and porcine peripheral nerve. Journal of Physiology, 513.2, 571585.Google Scholar
Maki, H. & Ilmoniemi, R. J. (2010). The relationship between peripheral and early cortical activation induced by transcranial magnetic stimulation. Neuroscience Letters, 478: 2428.Google Scholar
Mancini, M., Pellicciari, M. C., Brignani, D., Mauri, P., De Marchis, C., Miniussi, C., & Conforto, S. (2015). Automatic artifact suppression in simultaneous tDCS-EEG using adaptive filtering. Conference Proceedings IEEE Engineering, Medical and Biological Society.Google Scholar
Mangia, A. L., Pirini, M., & Cappello, A. (2014). Transcranial direct current stimulation and power spectral parameters: a tDCS/EEG co-registration study. Frontiers of Human Neuroscience, 8: 601.Google Scholar
Matthews, N., Luber, B., Qian, N., & Lisanby, S. (2001). Transcranial magnetic stimulation differentially affects speed and direction judgments. Experimental Brain Research, 140: 397406.Google Scholar
McConnell, K. A., Nahas, Z., Shastri, A., Lorberbaum, J. P., Kozel, F. A., Bohning, D. E., & George, M. S. (2001). The transcranial magnetic stimulation motor threshold depends on the distance from coil to underlying cortex: a replication in healthy adults comparing two methods of assessing the distance to cortex. Biological Psychiatry, 49: 454459.Google Scholar
McKinley, R. A., Bridges, N., Walters, C. M., & Nelson, J. (2012). Modulating the brain at work using noninvasive transcranial stimulation. NeuroImage, 59: 129137.Google Scholar
Mennemeier, M. S., Triggs, W., Chelette, K, C., Woods, A. J., Kimbrell, T., & Domhoffer, J. (2009). Sham transcranial magnetic stimulation using electrical stimulation of the scalp. Brain Stimulation, 2: 169173.Google Scholar
Meuller, J. K., Grigsby, E. M., Prevosto, V., Petraglia, F. W. III, Rao, H., Deng, Z.-D., … & Grill, W. M. (2014). Simultaneous transcranial magnetic stimulation and single-neuron recording in alert non-human primates. Nature Neuroscience, 17: 11301136.Google Scholar
Mills, K. R. & Nithi, K. A. (1997). Corticomotor threshold to magnetic stimulation: normal values and repeatability. Muscle Nerve, 20: 570576.Google Scholar
Miniussi, C., Harris, J. A., & Ruzzoli, M. (2013). Modeling non-invasive brain stimulation in cognitive neuroscience. Neuroscience & Biobehavioral Reviews, 37: 17021712.Google Scholar
Miniussi, C., Ruzzoli, M., & Walsh, V. (2010). The mechanism of transcranial magnetic stimulation in cognition. Cortex, 46: 128130.Google Scholar
Miranda, P. C., Hallett, M., & Basser, P. J. (2003). The electric field induced in the brain by magnetic stimulation: a 3-D finite element analysis of the effect of tissue heterogeneity and anisotropy. IEEE Transactions on Biomedical Engineering, 50: 10741085.Google Scholar
Mishory, A., Molnar, C., Koola, J., Li, X., Kozel, F. A., Myrick, H., … & George, M. S. (2004). The maximum-likelihood strategy for determining transcranial magnetic stimulation motor threshold, using parameter estimation by sequential testing is faster than conventional methods with similar precision. Journal of ECT, 20: 160165.Google Scholar
Moliadze, V., Giannikopoulos, D., Eysel, U. T., & Funke, K. (2005). Paired-pulse transcranial magnetic stimulation protocol applied to visual cortex of anaesthetized cat: effects on visually evoked single-unit activity. Journal of Physiology, 566: 955965.Google Scholar
Moliadze, V., Zhao, Y., Eysel, U., & Funke, K. (2003). Effect of transcranial magnetic stimulation on single-unit activity in the cat primary visual cortex. Journal of Physiology, 553: 665679.Google Scholar
Möller, C., Arai, N., Lücke, J., & Ziemann, U. (2009). Hysteresis effects on the input–output curve of motor evoked potentials. Clinical Neurophysiology, 120: 10031008.Google Scholar
Mottaghy, F. M., Gangitano, M., Sparing, R., & Pascual-Leone, A. (2002). Segregation of areas related to visual working memory in the prefrontal cortex revealed by rTMS. Cerebral Cortex, 12: 369375.Google Scholar
Muller-Dahlhaus, F. & Ziemann, U. (2015). Metaplasticity in human cortex. The Neuroscientist, 21: 185202.Google Scholar
Muller-Dahlhaus, F., Ziemann, U., & Classen, J. (2010). Plasticity resembling spike-timing dependent synaptic plasticity: the evidence in human cortex. Frontiers in Synaptic Neuroscience, 2: article 34.Google Scholar
Nagarajan, S. S., Durand, D. M., & Warman, E. N. (1993). Effects of induced electrical fields on finite neuronal structures: a simulation study. IEEE Transactions on Biomedical Engineering, 40: 11751188.Google Scholar
Nahas, Z., Li, X., Kozel, F. A., Mirzki, D., Memon, M., Miller, K., … & George, M. S. (2004). Safety and benefits of distance-adjusted prefrontal transcranial magnetic stimulation in depressed patients 55–75 years of age: a pilot study. Depression and Anxiety, 19: 249256.Google Scholar
National Research Council (1996). Possible Health Effects of Exposure to Residential Electric and Magnetic Fields. Washington, DC: National Academy Press.Google Scholar
Nitsche, M. A., Fricke, K., Henschke, U., Schlitterlau, A., Liebetanz, D., Lang, N., … & Paulus, W. (2003). Pharmacological modulation of cortical excitability shifts induced by transcranial direct current stimulation in humans. Journal of Physiology, 553: 293301.Google Scholar
Nitsche, M. A. & Paulus, W. (2000). Excitability changes induced in the human motor cortex by weak transcranial direct current stimulation. Journal of Physiology, 527: 633639.Google Scholar
Nitsche, M. A. & Paulus, W. (2001). Sustained excitability elevations induced by transcranial DC motor cortex stimulation in humans. Neurology, 57: 18991901.Google Scholar
Nitsche, M. A. & Paulus, W. (2011). Transcranial direct current stimulation: update 2011. Restorative Neurology and Neuroscience, 29: 463492.Google Scholar
Nitsche, M. A., Seeber, A., Frommann, K., Klein, C. C., Rochford, C., Nitsche, M. S., … & Tergau, F. (2005). Modulating parameters of excitability during and after transcranial direct current stimulation of the human motor cortex. Journal of Physiology, 568: 291303.Google Scholar
Ogiue-Ikeda, M., Kawato, S., & Ueno, S. (2003). The effect of transcranial magnetic stimulation on long-term potentiation in rat hippocampus. IEEE Transactions in Magnetics, 39: 3390.Google Scholar
Okabe, S., Ugawa, Y., & Kanazawa, I. (2003). 0.2-Hz repetitive transcranial magnetic stimulation has no add-on effects as compared to a realistic sham stimulation in Parkinson’s disease. Movement Disorders, 18: 382388.Google Scholar
Opitz, A., Legon, W., Rowlands, A., Bickel, W. K., Paulus, W., & Tyler, W. J. (2013). Physiological observations validate finite element models for estimating subject-specific electric field distributions induced by transcranial magnetic stimulation of the human motor cortex. NeuroImage, 81: 253–64.Google Scholar
O’Reardon, J. P., Solvason, H. B., Janicak, P. G., Sampson, S., Isenberg, K. E., Nahas, Z., … & Sackeim, H. A. (2007). Efficacy and safety of transcranial magnetic stimulation in the acute treatment of major depression: a multisite randomized controlled trial. Biological Psychiatry, 62: 12081216.Google Scholar
Parent, A. (2004). Giovanni Aldini: from animal electricity to human brain stimulation. Canadian Journal of Neurological Science, 31: 576584.Google Scholar
Parkin, B. L., Ekhtiari, H., & Walsh, V. F. (2015). Non-invasive human brain stimulation in cognitive neuroscience: a primer. Cell, 87: 932945.Google Scholar
Pascual-Leone, A., Cohen, L. G., Shotland, L. I., Dang, N., Pikus, A., Wassermann, E. M., … & Hallett, M. (1992). No evidence of hearing loss in humans due to transcranial magnetic stimulation. Neurology, 41: 647651.Google Scholar
Pascual-Leone, A., Gates, J. R., & Dhuna, A. (1991). Induction of speech arrest and counting errors with rapid-rate transcranial magnetic stimulation. Neurology, 41: 697701.Google Scholar
Pascual-Leone, A. & Hallett, M. (1994). Induction of errors in a delayed response task by repetitive transcranial magnetic stimulation of the dorsolateral prefrontal cortex. Neuroreport, 5: 25172520.Google Scholar
Pascual-Leone, A., Valls-Solé, J., Wassermann, E. M., & Hallett, M. (1994). Responses to rapid-rate transcranial magnetic stimulation of the human motor cortex. Brain, 117: 847858.Google Scholar
Pascual-Leone, A., Walsh, V., & Rothwell, J. (2000). Transcranial magnetic stimulation in cognitive neuroscience: virtual lesion, chronometry, and functional connectivity. Current Opinion in Neurobiology, 10: 232237.Google Scholar
Pasley, B. N., Allen, E., & Freeman, R. D. (2009). State-dependent variability of neuronal responses to transcranial magnetic stimulation of the visual cortex. Neuron, 62: 291303.Google Scholar
Paulus, W., Classen, J., Cohen, L. G., Large, C. H., DiLazzaro, V., Nitsche, M., … & Ziemann, U. (2008). State of the art: pharmacologic effects on cortical excitability measures tested by transcranial magnetic stimulation. Brain Stimulation, 1: 151163.Google Scholar
Paus, T. (2005). Inferring causality in brain images: a perturbation approach. Philosophical Transactions of the Royal Society B: Biological Sciences, 360: 11091114.Google Scholar
Paus, T., Jech, R., Thompson, C. J., Comeau, R., Peters, T., & Evans, A. C. (1997). Transcranial magnetic stimulation during positron emission tomography: a new method for studying connectivity of the human cerebral cortex. Journal of Neuroscience, 17: 31783184.Google Scholar
Peinemann, A., Reimer, B., Loer, C., Quartarone, A., Munchau, A., Conrad, B., & Siebner, H. R. (2004). Long-lasting increase in corticospinal excitability after 1800 pulses of subthreshold 5 Hz repetitive TMS to the primary motor cortex. Clinical Neurophysiology, 115: 15191526.Google Scholar
Pell, G. S., Roth, Y., & Zangen, A. (2011). Modulation of cortical excitability induced by repetitive transcranial magnetic stimulation: influence of timing and geometrical parameters and underlying mechanisms. Progress in Neurobiology, 93: 5998.Google Scholar
Peterchev, A. V., Deng, Z.-D., & Goetz, S. M. (2015). Advances in transcranial magnetic stimulation technology. In Reti, I. M. (ed.), Brain Stimulation: Methodologies and Interventions (pp. 165189). Hoboken, NJ: Wiley-Blackwell.Google Scholar
Peterchev, A. V., Jalinous, R., & Lisanby, S. H. (2008). A transcranial magnetic stimulator inducing near-rectangular pulses with controllable pulse width (cTMS). IEEE Transactions in Biomedical Engineering, 55: 257266.Google Scholar
Pitcher, D., Garrido, L., Duchaine, B., & Walsh, V. (2008). Transcranial magnetic stimulation disrupts the perception and embodiment of facial expressions. Journal of Neuroscience, 28: 89298933.Google Scholar
Pitcher, D., Goldhaber, T., Duchaine, B., Walsh, V., & Kanwisher, N. (2012). Two critical and functionally distinct stages of face and body perception. Journal of Neuroscience, 32: 1587715885.Google Scholar
Pitcher, J. B., Ogston, K. M., & Miles, T. S. (2003). Age and sex differences in human motor cortex input–output characteristics. Journal of Physiology, 546: 605613.Google Scholar
Plow, E. B., Cattaneo, Z., Carlson, T. A., Alvarez, G. A., Pascual-Leone, A., & Battelli, L. (2014). The compensatory dynamic of inter-hemispheric interactions in visuospatial attention revealed using rTMS and fMRI. Frontiers in Human Neuroscience, 8: 226.Google Scholar
Poreisz, C., Boros, K., Antal, A., & Paulus, W. (2007). Safety aspects of transcranial direct current stimulation concerning healthy subjects and patients. Brain Research Bulletin, 72: 208214.Google Scholar
Qi, F., Wu, A. D., & Schweighofer, N. (2011). Fast estimation of transcranial magnetic stimulation motor threshold. Brain Stimulation, 4: 5057.Google Scholar
Radman, T., Ramos, R. L., Brumberg, J. C., & Bikson, M. (2009). Role of cortical cell type and morphology in subthreshold and suprathreshold uniform electric field stimulation in vitro. Brain Stimulation, 2: 215228.Google Scholar
Ragert, P., Dinse, H. R., Pleger, B., Wilimzig, C., Frombach, E., Schwenkreis, P., & Tegenthoff, M. (2003). Combination of 5 Hz repetitive transcranial stimulation (rTMS) and tactile coactivation boosts tactile discrimination in humans. Neuroscience Letters, 348: 105108.Google Scholar
Rahman, A., Reato, D., Arlotti, M., Gasca, F., Datta, A., Parra, L.C., & Bikson, M. (2013). Cellular effects of acute direct current stimulation: somatic and synaptic terminal effects. Journal of Physiology, 591: 25632578.Google Scholar
Rahnev, D. A., Maniscalco, B., Luber, B., Lau, H., & Lisanby, S. H. (2011). Direct injection of noise to the visual cortex decreases accuracy but increases decision confidence. Journal of Neurophysiology, 107: 15561563.Google Scholar
Rauschecker, A. M., Bestman, S., Walsh, V., & Thilo, K. V. (2004). Phosphene threshold as a function of contrast of external visual stimuli. Experimental Brain Research, 157: 124127.Google Scholar
Ray, P. G., Meador, K. J., Epstein, C. M., Loring, D. W., & Day, L. J. (1998). Magnetic stimulation of the visual cortex: factors influencing the perception of phosphenes. Journal of Clinical Neurophysiology, 15: 351357.Google Scholar
Reid, A. E., Chiappa, K. H., & Cros, D. (2002). Motor threshold, facillitation and the silent period in cortical magnetic stimulation. In Pascual-Leone, A., Davey, N. J., Rothwell, J., Wassermann, E. M., & Puri, B. K. (eds.), Handbook of Transcranial Magnetic Stimulation (pp. 9711). London: Arnold.Google Scholar
Romei, V., Brodbeck, V., Michel, C., Amedi, A., Pascual-Leone, A., & Thut, G. (2008) Spontaneous fluctuations in posterior alpha-band EEG activity reflect variability in excitability of human visual areas. Cerebral Cortex, 18: 20102018.Google Scholar
Romei, V., Driver, J., Schyns, P. G., & Thut, G. (2011). Rhythmic TMS over parietal cortex links distinct brain frequencies to global versus local visual processing. Current Biology, 21: 334337.Google Scholar
Romei, V., Gross, J., & Thut, G. (2010). On the role of prestimulus alpha rhythms over occipito-parietal areas in visual input regulation: correlation or causation? Journal of Neuroscience, 30: 86928697.Google Scholar
Romeo, S., Gileo, F., Pedace, F., Ozkaynak, S., Inghilleri, M., Manfredi, M., & Berardelli, A. (2000). Changes in the cortical silent period after repetitive magnetic stimulation of cortical motor areas. Experimental Brain Research, 135: 504510.Google Scholar
Rossi, S., Hallett, M., Rossini, P. M., Pascual-Leone, A., & The Safety of TMS Consensus Group (2009). Safety, ethical considerations, and application guidelines for the use of transcranial magnetic stimulation in clinical practice and research. Clinical Neurophysiology, 120: 20082039.Google Scholar
Rossini, P. M., Barker, A. T., Berardelli, A., Caramia, M. D., Caruso, G., Cracco, R. Q., … & Tomberg, C. (1994). Non-invasive electrical and magnetic stimulation of the brain, spinal cord and roots: basic principles and procedures for routine clinical application. Report of an IFCN committee. Electroencephalography & Clinical Neurophysiology, 91: 7992.Google Scholar
Roth, B. J., Pascual-Leone, A., Cohen, L. G., & Hallett, M. (1992). The heating of metal electrodes during rapid-rate magnetic stimulation: a possible safety hazard. Electroencephalography & Clinical Neurophysiology, 85: 116123.Google Scholar
Rothkegel, H., Sommer, M., & Paulus, W. (2010). Breaks during 5 Hz rTMS are essential for facilitatory after effects. Clinical Neurophysiology, 121: 426430.Google Scholar
Rothwell, J. C., Hallett, M., Berardelli, A., Eisen, A., Rossini, P., & Paulus, W. (1999). Magnetic stimulation: motor evoked potentials. The International Federation of Clinical Neurophysiology. Electroencephalography & Clinical Neurophysiology Supplement, 52: 97103.Google Scholar
Roy, A., Baxter, B., & He, B. (2014). High-definition transcranial direct current stimulation induces both acute and persistent changes in broadband cortical synchronization: a simultaneous tDCS-EEG study. IEEE Transactions in Biomedical Engineering, 61: 19671978.Google Scholar
Rudiak, D. & Marg, E. (1994). Finding the depth of magnetic brain stimulation: a re-evaluation. Electroencephalography & Clinical Neurophysiology, 93: 358371.Google Scholar
Ruff, C. C., Bestmann, S., Blankenburg, F., Bjoertomt, O., Josephs, O., Weiskopf, N., … & Driver, J. (2008). Distinct causal influences of parietal versus frontal areas on human visual cortex: evidence from concurrent TMS–fMRI. Cerebral Cortex, 18: 817827.Google Scholar
Ruff, C. C., Blankenburg, F., Bjoertomt, O., Bestmann, S., Freeman, E., Haynes, J. D., … & Driver, J. (2006). Concurrent TMS–fMRI and psychophysics reveal frontal influences on human retinotopic visual cortex. Current Biology, 16: 14791488.Google Scholar
Ruohonen, J., Ollikainen, M., Nikouline, V., Virtanen, J., & Ilmoniemi, R. (2000). Coil design for real and sham transcranial magnetic stimulation. IEEE Transactions on Biomedical Engineering, 47: 145148.Google Scholar
Sack, A. T. (2010). Does TMS need functional imaging? Cortex, 46: 131133.Google Scholar
Sack, A. T., Cohen Kadosh, R., Schuhmann, T., Moerel, M., Walsh, V., & Goebel, R. (2009). Optimizing functional accuracy of TMS in cognitive studies: a comparison of methods. Journal of Cognitive Neuroscience, 21: 207221.Google Scholar
Sakkas, P., Theleritis, C. G., Psarros, C., Papadimitriou, G. N., & Soldatos, C. R. (2008). Jacksonian seizure in a manic patient treated with rTMS. World Journal of Biological Psychiatry, 9: 159160.Google Scholar
Sale, M. V., Ridding, M. C., & Nordstrom, M. A. (2007). Factors influencing the magnitude and reproducibility of corticomotor excitability changes induced by paired associative stimulation. Experimental Brain Research, 181: 615626.Google Scholar
Schnitzler, A. & Gross, J. (2005). Normal and pathological oscillatory communication in the brain. Nature Reviews Neuroscience, 6: 285296.Google Scholar
Schrader, L. M., Stern, J. M., Koski, L., Nuwer, M. R., & Engel, J. Jr. (2004). Seizure incidence during single- and paired-pulse transcranial magnetic stimulation (TMS) in individuals with epilepsy. Clinical Neurophysiology, 115: 27282737.Google Scholar
Schwarzkopf, D. S., Silvanto, J., & Rees, G. (2011). Stochastic resonance effects reveal the neural mechanisms of transcranial magnetic stimulation. Journal of Neuroscience, 31: 31433147.Google Scholar
Shastri, A., George, M. S., & Bohning, D. E. (1999). Performance of a system for interleaving transcranial magnetic stimulation with steady-state magnetic resonance imaging. Electroencephalography & Clinical Neurophysiology Supplement, 51: 5564.Google Scholar
Siebner, H. R., Hartwigsen, G., Kassuba, T., & Rothwell, J. C. (2009). How does transcranial magnetic stimulation modify neuronal activity in the brain? Implications for studies of cognition. Cortex, 45: 10351042.Google Scholar
Siebner, H. R., Lang, N., Rizzo, V., Nitsche, M. A., Paulus, W., Lemon, R. N., & Rothwell, J. C. (2004). Preconditioning of low-frequency repetitive transcranial magnetic stimulation with transcranial direct current stimulation: evidence for homeostatic plasticity in the human motor cortex. Journal of Neuroscience, 24: 33793385.Google Scholar
Silva, S., Basser, P. J., & Miranda, P. C. (2008). Elucidating the mechanisms and loci of neuronal excitation by transcranial magnetic stimulation using a finite element model of a cortical sulcus. Clinical Neurophysiology, 119: 24052413.Google Scholar
Silvanto, J., Cattaneo, Z., Battelli, L., & Pascual-Leone, A. (2008a). Baseline cortical excitability determines whether TMS disrupts or facilitates behavior. Journal of Neurophysiology, 99: 27252730.Google Scholar
Silvanto, J., Lavie, N., & Walsh, V. (2006). Stimulation of the human frontal eye fields modulates sensitivity of extrastriate visual cortex. Journal of Neurophysiology, 96: 941945.Google Scholar
Silvanto, J., Muggleton, N., & Walsh, V. (2008b). State-dependency in brain stimulation studies of perception and cognition. Trends in Cognitive Sciences, 12: 447454.Google Scholar
Silvanto, J. & Pascual-Leone, A. (2012). Why the assessment of causality in brain behavior relations requires brain stimulation. Journal of Cognitive Neuroscience, 24: 775777.Google Scholar
Sirota, A., Montgomery, S., Fujisawa, S., Isomura, Y., Zugaro, M., & Buzsaki, G. (2009). Entrainment of neocortical neurons and gamma oscillations by the hippocampal theta rhythm. Neuron, 60: 683697.Google Scholar
Sommer, J., Jansen, A., Dräger, B., Steinsträter, O., Breitenstein, C., Deppe, M., & Knecht, S. (2006). Transcranial magnetic stimulation: a sandwich coil design for a better sham. Clinical Neurophysiology, 117: 440446.Google Scholar
Sommer, M., Lang, N., Tergau, F., & Paulus, W. (2002). Neuronal tissue polarization induced by repetitive transcranial magnetic stimulation? Neuroreport, 13: 809811.Google Scholar
Sparing, R., Hesse, M. D., & Fink, G. R. (2010). Neuronavigation for transcranial magnetic stimulation (TMS): where we are and where we are going. Cortex, 46: 118120.Google Scholar
Speer, A. M, Kimbrell, T. A., Wassermann, E. M., Repella, J. D., Willis, M. W., Herscovitch, P., & Post, R. M. (2000). Opposite effects of high and low frequency rTMS on regional brain activity in depressed patients. Biological Psychiatry, 48: 11331141.Google Scholar
Stefan, K., Kunesch, E., Benecke, R., Cohen, L. G., & Classen, J. (2002). Mechanisms of enhancement of human motor cortex excitability induced by interventional paired associative stimulation. Journal of Physiology, 543: 699708.Google Scholar
Stefan, K., Kunesch, E., Cohen, L. G., Benecke, R., & Classen, J. (2000). Induction of plasticity in the human motor cortex by paired associative stimulation. Brain, 123: 572584.Google Scholar
Stewart, L. M., Walsh, V., & Rothwell, J. C. (2001). Motor and phosphene thresholds: a transcranial magnetic stimulation correlation study. Neuropsychologia, 39: 415419.Google Scholar
Stinear, C. M. & Byblow, W. D. (2003). Motor imagery of phasic thumb abduction temporally and spatially modulates corticospinal excitability. Clinical Neurophysiology, 114: 909914.Google Scholar
Strafella, A. P. & Paus, T. (2000). Modulation of cortical excitability during action observation: a transcranial magnetic stimulation study. Neuroreport, 11: 22892292.Google Scholar
Taylor, J. L. & Loo, C. K. (2007). Stimulus waveform influences the efficacy of repetitive transcranial magnetic stimulation. Journal of Affective Disorders, 97: 271276.Google Scholar
Terao, Y. & Ugawa, Y. (2002). Basic mechanisms of TMS. Journal of Clinical Neurophysiology, 19: 322343.Google Scholar
Terao, Y., Ugawa, Y., Suzuki, M., Sakai, K., Hanajima, R., Gemba-Shimizu, K., & Kanazawa, I. (1997). Shortening of simple reaction time by peripheral electrical and submotor-threshold magnetic cortical stimulation. Experimental Brain Research, 115: 541545.Google Scholar
Thickbroom, G. W. (2007). Transcranial magnetic stimulation and synaptic plasticity: experimental framework and human models. Experimental Brain Research, 180: 583593.Google Scholar
Thielscher, A., Opitz, A., & Windhoff, M. (2011). Impact of the gyral geometry on the electric field induced by transcranial magnetic stimulation. NeuroImage, 54: 234243.Google Scholar
Thut, G. & Miniussi, C. (2009). New insights into rhythmic brain activity from TMS–EEG studies. Trends in Cognitive Sciences, 13: 182189.Google Scholar
Thut, G., Miniussi, C., & Gross, J. (2012). The functional importance of rhythmic activity in the brain. Current Biology, 22: R658R663.Google Scholar
Thut, G. & Pascual-Leone, A. (2009). A review of combined TMS-EEG studies to characterize lasting effects of repetitive TMS and assess their usefulness in cognitive and clinical neuroscience. Brain Topography, 22: 219232.Google Scholar
Thut, G., Schyns, P. G., & Gross, J. (2011a). Entrainment of perceptually relevant brain oscillations by non-invasive rhythmic stimulation of the human brain. Frontiers in Psychology, 2: 170.Google Scholar
Thut, G., Veniero, D., Romei, V., Miniussi, C., Schyns, P., & Gross, J. (2011b). Rhythmic TMS causes local entrainment of natural oscillatory signatures. Current Biology, 21: 11761185.Google Scholar
Tischler, H., Wolfus, S., Friedman, A., Perel, E., Pashut, T., Lavidor, M., … & Bar-Gad, I. (2011). Mini-coil for magnetic stimulation in the behaving primate. Journal of Neuroscience Methods, 194: 242251.Google Scholar
Touge, T., Gerschlager, W., Brown, P., & Rothwell, J. C. (2001). Are the after-effects of low-frequency rTMS on motor cortex excitability due to changes in the efficacy of cortical synapses? Clinical Neurophysiology, 112: 21382145.Google Scholar
Tranulis, C., Guéguen, B., Pham-Scottez, A., Vacheron, M., Cabelguen, G., Costantini, A., … & Galinovski, A.(2006). Motor threshold in transcranial magnetic stimulation: comparison of three estimation methods. Clinical Neurophysiology, 36: 17.Google Scholar
Valls-Sole, J., Pascual-Leone, A., Wassermann, E. M., & Hallett, M. (1992). Human motor evoked responses to paired transcranial magnetic stimuli. Electroencephalography & Clinical Neurophysiology, 85: 355364.Google Scholar
Van Der Werf, Y. D. & Paus, T. (2006). The neural response to transcranial magnetic stimulation of the human motor cortex. I: Intracortical and cortico-cortical contributions. Experimental Brain Research, 175: 231245.Google Scholar
Virtanen, J., Ruohonen, J., Näätänen, R., & Ilmoniemi, R. J. (1999). Instrumentation for the measurement of electric brain responses to transcranial magnetic stimulation. Medical Biological Engineering and Computing, 37: 322326.Google Scholar
Wagner, T., Valero-Cabré, A., & Pascual-Leone, A. (2007). Noninvasive brain stimulation. Annual Review of Biomedical Engineering, 9: 527565.Google Scholar
Walsh, V., Ellison, A., Battelli, L., & Cowey, A. (1998). Task-specific impairments and enhancements induced by magnetic stimulation of human visual area V5. Proceedings in the Biological Sciences, 265: 537543.Google Scholar
Walsh, V. & Pascual-Leone, A. (2003). Transcranial Magnetic Stimulation: A Neurochronometrics of Mind. Vol. 1. Boston, MA: MIT Press.Google Scholar
Walsh, V. & Rushworth, M. (1999). A primer of magnetic stimulation as a tool for neuropsychology. Neuropsychologia, 37: 125135.Google Scholar
Wang, J. X., Rogers, L. M., Gross, E. Z., Ryals, A. J., Dokucu, M. E., Brandstatt, K. I., … & Voss, J. I. (2014). Targeted enhancement of cortical-hippocampal brain networks and associative memory. Science, 345: 10541057.Google Scholar
Wassermann, E. M. (1998). Risk and safety of repetitive transcranial magnetic stimulation. Electroencephalography & Clinical Neurophysiology, 108: 116.Google Scholar
Watanabe, T., Hanajima, R., Shirota, Y., Ohminami, S., Tsutsumi, R., Terao, Y., … & Ohtomo, K. (2014). Bidirectional effects on interhemispheric resting-state functional connectivity induced by excitatory and inhibitory repetitive transcranial magnetic stimulation. Human Brain Mapping, 35: 18961905.Google Scholar
Weiss, C., Nettekoven, C., Rehme, A. K., Neuschmelting, V., Eisenbeis, A., Goldbrunner, R., & Grefkes, C. (2013). Mapping the hand, foot and face representations in the primary motor cortex: retest reliability of neuronavigated TMS versus functional MRI. NeuroImage, 66: 531542.Google Scholar
Weissman, J. D., Epstein, C. M., & Davey, K. R. (1992). Magnetic brain stimulation and brain size: relevance to animal studies. Electroencephalography & Clinical Neurophysiology, 85: 215219.Google Scholar
Weisz, N., Steidle, L., & Lorenz, I. (2012). Formerly known as inhibitory: effects of 1 Hz rTMS on auditory cortex are state-dependent. European Journal of Neuroscience, 36: 20772087.Google Scholar
Werhahn, K. J., Kunesch, E., Noachtar, S., Benecke, R., & Classen, J. (1999). Differential effects on motorcortical inhibition induced by blockade of GABA uptake in humans. Journal of Physiology, 517: 591597.Google Scholar
Westin, G. G., Bassi, B. D., Lisanby, S. H., & Luber, B. (2014). Determination of motor threshold using visual observation overestimates transcranial magnetic stimulation dosage: safety implications. Clinical Neurophysiology, 125: 142147.Google Scholar
Windhoff, M., Opitz, A., & Thielscher, A. (2013). Electric field calculations in brain stimulation based on finite elements: an optimized processing pipeline for the generation and usage of accurate individual head models. Human Brain Mapping, 34: 923935.Google Scholar
Wipfli, M., Felblinger, J., Mosimann, U. P., Hess, C. W., Schlaepfer, T. E., & Muri, R. M. (2001). Double-pulse transcranial magnetic stimulation over the frontal eye field facilitates triggering of memory-guided saccades. European Journal of Neuroscience, 14: 571575.Google Scholar
Wischnewski, M. & Schutter, D. J. L. G. (2015). Efficacy and time course of theta burst stimulation in healthy humans. Brain Stimulation, 8: 685692.Google Scholar
Wolters, A., Sandbrink, F., & Schlottmann, A. (2003). A temporally asymmetric Hebbian rule governing plasticity in the human motor cortex. Journal of Neurophysiology, 89: 23392345.Google Scholar
Yamanaka, K., Yamagata, B., Tomioka, H., Kawasaki, S., & Mimura, M. (2010). Transcranial magnetic stimulation of the parietal cortex facilitates spatial working memory: near infrared spectroscopy study. Cerebral Cortex, 20: 10371045.Google Scholar
Zangen, A., Roth, Y., Voller, B., & Hallett, M. (2005). Transcranial magnetic stimulation of deep brain regions: evidence for efficacy of the H-coil. Clinical Neurophysiology, 116: 775779.Google Scholar
Ziemann, U. (2002). Paired pulse techniques. In Pascual-Leone, A., Davey, N. J., Rothwell, J., Wassermann, E. M., & Puri, B. K. (eds.), Handbook of Transcranial Magnetic Stimulation, vol. 1 (pp. 141159). London: Arnold.Google Scholar
Ziemann, U. (2010). TMS in cognitive neuroscience: virtual lesion and beyond. Cortex, 46: 124127.Google Scholar
Ziemann, U. (2011). Transcranial magnetic stimulation at the interface with other techniques: a powerful tool for studying the human cortex. The Neuroscientist, 17: 368381.Google Scholar
Ziemann, U., Lonnecker, S., Steinhoff, B. J., & Paulus, W. (1996a). The effect of lorazepam on the motor cortical excitability in man. Experimental Brain Research, 109: 127135.Google Scholar
Ziemann, U., Paulus, W., Nitsche, M. A. Pascual-Leone, A., Byblow, W. D., Berardelli, A., … & Rothwell, J. C. (2008). Consensus: motor cortex plasticity protocols. Brain Stimulation, 1: 164182.Google Scholar
Ziemann, U. & Rothwell, J. C. (2000). I-waves in motor cortex. Journal of Clinical Neurophysiology, 17: 397405.Google Scholar
Ziemann, U., Rothwell, J. C., & Ridding, M. C. (1996b). Interaction between intracortical inhibition and facilitation in human motor cortex. Journal of Physiology, 496: 873881.Google Scholar

References

Abbs, J. H., Gracco, V. L., & Blair, C. (1984). Functional muscle partitioning during voluntary movement: facial muscle activity for speech. Experimental Neurology, 85: 469479.Google Scholar
Adrian, E. D. & Bronk, D. W. (1929). The discharge of impulses in motor nerve fibers. Part II: the frequency of discharge in reflex and voluntary contractions. Journal of Physiology, 67: 119151.Google Scholar
Alaoui-Ismaïli, O., Vernet-Maury, E., Dittmar, A., Delhomme, G., & Chanel, J. (1997). Odor hedonics: connection with emotional response estimated by autonomic parameters. Chemical Senses, 22: 237248.Google Scholar
Alexander, A. B. & Smith, D. D. (1979). Clinical applications of EMG biofeedback. In Gatchel, R. I. & Price, K. P. (eds.), Clinical Applications of Biofeedback: Appraisal and Status (pp. 112133). New York: Pergamon.Google Scholar
Alius, M. G., Pane-Farre, C. A., Low, A., & Hamm, A. O. (2015). Modulation of the blink reflex and P3 component of the startle response during an interoceptive challenge. Psychophysiology, 52: 140148.Google Scholar
Allain, S., Carbonnell, L., Burle, B., Hasbroucq, T., & Vidal, F. (2004). On-line executive control: an electromyographic study. Psychophysiology, 41: 113116.Google Scholar
Allport, G. W. (1968). The historical background of modern social psychology. In Lindzey, G. & Aronson, E. (eds.), The Handbook of Social Psychology, 2nd edn. Menlo Park, CA: Addison-Wesley.Google Scholar
Anthony, B. I. (1985). In the blink of an eye: implications of reflex modifcation for information processing. In Ackles, P. K., Jennings, J. R., & Coles, M. G. H. (eds.), Advances in Psychophysiology, vol. 1 (pp. 167218). Greenwich, CT: JAI Press.Google Scholar
Arndt, J., Allen, J. J. B., & Greenberg, J. (2001). Traces of terror: subliminal death primes and facial electromyographic indices of affect. Motivation and Emotion, 25: 253277.Google Scholar
Bakker, F. C., Boschker, M. S. J., & Chung, T. (1996). Changes in muscular activity while imagining weight lifting using stimulus or response propositions. Journal of Sport and Exercise Psychology, 18: 313324.Google Scholar
Bansevicius, D. & Sjaastad, O. (1996). Cervicogenic headache: the influence of mental load on pain level and EMG of the shoulder-neck and facial muscles. Headache, 36: 372378.Google Scholar
Bartholow, B. D., Fabiani, M., Gratton, G., & Bettencourt, B. A. (2001). A psychophysiological examination of cognitive processing of affective responses to social expectancy violations. Psychological Science, 12: 197204.Google Scholar
Basmajian, J. V. (1989). Biofeedback: Principles and Practice for Clinicians, 3rd edn. Baltimore, MD: Williams & Wilkins.Google Scholar
Basmajian, J. V. & De Luca, C. J. (1985). Muscles Alive: Their Functions Revealed by Electromyography, 5th edn. Baltimore, MD: Williams & Wilkins.Google Scholar
Bavelas, J. B., Black, A., Chovil, N., Lemery, C. R., & Mullett, J. (1988). Form and function in motor mimicry: topographic evidence that the primary function is communicative. Human Communication Research, 14: 275299.Google Scholar
Bavelas, J. B., Black, A., Lemery, C. R., & Mullett, J. (1986). “I show how you feel”: motor mimicry as a communicative act. Journal of Personality and Social Psychology, 50: 322329.Google Scholar
Beebe-Center, J. G. (1932). The Psychology of Pleasantness and Unpleasantness. New York: Van Nostrand.Google Scholar
Berkinblit, M. B., Feldman, A. G., & Fulson, O. I. (1986). Adaptability of innate motor patterns and motor control mechanisms. Behavioral and Brain Sciences, 9: 585638.Google Scholar
Bernardis, P. & Gentilucci, M. (2006). Speech and gesture share the same communication system. Neuropsychologia, 44: 178190.Google Scholar
Blumenthal, T., Cuthbert, B. N., Filion, D. L., Hackley, S., Lipp, O. V., & van Boxtel, A. (2005). Committee report: guidelines for human startle eyeblink electromyographic studies. Psychophysiology, 42: 115.Google Scholar
Boiten, F. (1996). Autonomic response patterns during voluntary facial action. Psychophysiology, 33: 123131.Google Scholar
Bond, C. F. & Titus, L. J. (1983). Social facilitation: a meta-analysis of 241 studies. Psychological Bulletin, 94: 265292.Google Scholar
Braathen, E. T. & Sveback, S. (1994). EMG response patterns and motivational styles as predictors of performance and discontinuation in explosive and endurance sports among talented teenage athletes. Personality and Individual Differences, 17: 545556.Google Scholar
Bradley, M., Cuthbert, B. N., & Lang, P. J. (1999). Affect and the startle reflex. In Dawson, M. E., Schell, A., & Boehmelt, A. (eds.), Startle Modification: Implications for Neuroscience, Cognitive Science, and Clinical Science (pp. 157183). Cambridge University Press.Google Scholar
Breazeal, C. (2003). Emotion and sociable humanoid robots. International Journal of Human-Computer Studies, 59: 119155.Google Scholar
Britt, T. W. & Blumenthal, T. D. (1992). The effects of anxiety on motoric expression of the startle response. Personality and Individual Differences, 13: 9197.Google Scholar
Brodie, M., Walmsley, A., & Page, W. (2008). Fusion motion capture: a prototype system using inertial measurement units and GPS for the biomechanical analysis of ski racing. Sports Technology, 1: 1728.Google Scholar
Brown, P. (2000). Cortical drives to human muscles: the Piper and related rhythms. Progress in Neurobiology, 60: 97108.Google Scholar
Bruintjes, T. D., van Olphen, A. F., Hillen, B., & Weijs, W. A. (1996). Electromyography of the human nasal muscles. European Archives of Otorhinolaryngology, 253: 464469.Google Scholar
Brunia, C. H. M. & Boelhouwer, A. J. W. (1988). Reflexes as tools: a window in the central nervous system. Advances in Psychophysiology, 3: 167.Google Scholar
Budzynski, T. H. & Stoyva, I. M. (1969). An instrument for producing deep muscle relaxation by means of analog information feedback. Journal of Applied Behavior Analysis, 2: 231237.Google Scholar
Burtt, H. E. & Tuttle, W. W. (1925). The patellar tendon reflex and affective tone. American Journal of Psychology, 36: 553561.Google Scholar
Butko, N. J., Theocharous, G., Philipose, M., & Movellan, J. R. (2011). Automated facial affect analysis for one-on-one tutoring applications. In Proceedings of the 2011 IEEE International Conference on Automatic Face & Gesture Recognition and Workshops (FG 2011) (pp. 382387). Piscataway, NJ: IEEE.Google Scholar
Cacioppo, J. T., Berntson, G., & Klein, D. J. (1992a). What is an emotion? The role of somatovisceral afference, with special emphasis on somatovisceral “illusions.” Review of Personality and Social Psychology, 14: 6398.Google Scholar
Cacioppo, J. T., Bush, L. K., & Tassinary, L. G. (1992b). Microexpressive facial actions as a function of affective stimuli: replication and extension. Personality and Social Psychology Bulletin, 18: 515526.Google Scholar
Cacioppo, J. T., Martzke, J. S., Petty, R. E., & Tassinary, L. G. (1988). Specific forms of facial EMG response index emotions during an interview: from Darwin to the continuous flow hypothesis of affect-laden information processing. Journal of Personality and Social Psychology, 54: 592604.Google Scholar
Cacioppo, J. T. & Petty, R. E. (1981a). Electromyograms as measures of extent and affectivity of information processing. American Psychologist, 36: 441456.Google Scholar
Cacioppo, J. T. & Petty, R. E. (1981b). Electromyographic specificity during covert information processing. Psychophysiology, 18: 518523.Google Scholar
Cacioppo, J. T. & Petty, R. E. (1986). Social processes. In Coles, M. G. H., Donchin, E., & Porges, S. (eds.), Psychophysiology: Systems, Processes, and Applications (pp. 646679). New York: Guilford Press.Google Scholar
Cacioppo, J. T., Petty, R. E., Losch, M. E., & Kim, H. S. (1986). Electromyographic activity over facial muscle regions can differentiate the valence and intensity of affective reactions. Journal of Personality and Social Psychology, 50: 260268.Google Scholar
Cacioppo, J. T, Petty, R. E., & Morris, K. (1985). Semantic, evaluative, and self-referent processing: memory, cognitive effort, and somatovisceral activity. Psychophysiology, 22: 371384.Google Scholar
Cacioppo, J. T., Priester, J. T., & Berntson, G. (1993). Rudimentary determinants of attitudes: II. Arm flexion and extension have differential effects on attitudes. Journal of Personality and Social Psychology, 65: 517.Google Scholar
Cacioppo, J. T., Rourke, P. A., Marshall-Goodell, B. S., Tassinary, L. G., & Baron, R. S. (1990). Rudimentary physiological effects of mere observation. Psychophysiology, 27: 177186.Google Scholar
Cacioppo, J. T., Tassinary, L. G., & Fridlund, A. (1990). The skeletomotor system. In Cacioppo, J. T. & Tassinary, L. G. (eds.), Principles of Psychophysiology (pp. 325384). Cambridge University Press.Google Scholar
Cappella, J. N. (1993). The facial feedback hypothesis in human interaction: review and speculation. Journal of Language and Social Psychology, 12: 1329.Google Scholar
Castellano, G., Pereira, A., Leite, I., Paiva, A., & McOwan, P. W. (2009). Detecting user engagement with a robot companion using task and social interaction-based features. In Proceedings of the 2009 International Conference on Multimodal Interfaces (pp. 119126). New York: Association for Computing Machinery.Google Scholar
Cattaneo, L., Fabbri-Destro, M., Boria, S., Pieraccini, C., Monti, A., Cossu, G., & Rizzolatti, G. (2007). Impairment of actions chains in autism and its possible role in intention understanding. Proceedings of the National Academy of Sciences of the USA, 104: 1782517830.Google Scholar
Chang, K., Lu, S., & Wu, X. (2012). A wireless sEMG recording system and its application to muscle fatigue detection. Sensors, 12: 489499.Google Scholar
Chapman, A. J. (1974). An electromyographic study of social facilitation: a test of the “mere presence” hypothesis. British Journal of Psychology, 65: 123128.Google Scholar
Charbonnier, C., Kolo-Christophe, F., Duc, S., Pfirrmann, C., Menetrey, J., Duthon, V., … & Hoffmeyer, P. (2009). Extreme motion as a potential initiator of hip osteoarthritis. Swiss Medical Weekly Supplement, 173: 2324.Google Scholar
Chen, Y., Yang, Z., & Wang, J. (2015). Eyebrow emotional expression recognition using surface EMG signals. Neurocomputing, 168: 871879.Google Scholar
Chovil, N. (1991). Social determinants of facial displays. Journal of Nonverbal Behavior, 15: 141154.Google Scholar
Chun, W. X. (1985). An approach to the nature of tension headache. Headache, 25: 188189.Google Scholar
Cikara, M. & Fiske, S. T. (2012). Stereotypes and Schadenfreude: affective and physiological markers of pleasure at outgroup misfortunes. Social Psychological and Personality Science, 3: 6371.Google Scholar
Cloete, T. & Scheffer, C. (2010). Repeatability of an off-the-shelf, full body inertial motion capture system during clinical gait analysis. In Proceedings of the 2010 Annual International Conference of the IEEE on Engineering in Medicine and Biology Society (EMBC) (pp. 51255128). Piscataway, NJ: IEEE.Google Scholar
Compton, R. W. (1973). Morphological, physiological, and behavioral studies of the facial musculature of the Coati (Nasua). Brain, Behavior, and Evolution, 7: 85126.Google Scholar
Contarino, M. F., Groot, P. F. C., van der Meer, J. N., Bour, L. J., Speelman, J. D., Nederveen, , … & van Rootselaar, A. (2012). Is there a role for combined EMG–fMRI in exploring the pathophysiology of essential tremor and improving functional neurosurgery? PLoS One, 7: e46234. doi:10.1371/journal.pone.0046234Google Scholar
Cook, T. D. & Campbell, D. T. (1979). Quasi-Experimentation: Design and Analysis Issues for Field Settings. Boston, MA: Houghton Mifflin.Google Scholar
Cooper, R., Osselton, J. W., & Shaw, J. C. (1980). EEG Technology, 3rd edn. London: Butterworth.Google Scholar
Cram, J. R., Kasman, G., & Holtz, J. (1998). Introduction to Surface EMG. Gaithersburg, MD: Aspen Publications.Google Scholar
Cuthbertson, R. A. (1990). The highly original Dr. Duchenne. In Cuthbertson, R. A. (ed. and trans.), The Mechanism of Human Facial Expression (pp. 225241). Cambridge University Press.Google Scholar
Dai, Y., Shibata, Y., Ishii, T., Hashimoto, K., Katamachi, K., Noguchi, K., … & Cai, D. (2001). An associate memory model of facial expressions and its application in facial expression recognition of patients on bed. In Proceedings of the IEEE International Conference on Multimedia (p. 151). Piscataway, NJ: IEEE.Google Scholar
Dambrun, M., Despres, G., & Guimond, S. (2004). On the multifaceted nature of prejudice: psychophysiology responses to ingroup and outgroup ethnic stimuli. Current Research in Social Psychology, 8: 187204.Google Scholar
Darwin, C. (1873 [1872]). The Expression of the Emotions in Man and Animals. New York: Appleton.Google Scholar
Das, S., Trutoiu, L., Murai, A., Alcindor, D., Oh, M., De la Torre, F., & Hodgins, J. (2011). Quantitative measurement of motor symptoms in Parkinson’s disease: a study with full-body motion capture data. In Proceedings of the Engineering in Medicine and Biology Society, EMBC, 2011 Annual International Conference of the IEEE (pp. 67896792). Piscataway, NJ: IEEE.Google Scholar
Dauvilliers, Y., Rompré, S., Gagnon, J. F., Vendette, M., Petit, D., & Montplaisir, J. (2007). REM sleep characteristics in narcolepsy and REM sleep behavior disorder. Sleep, 30: 844849.Google Scholar
Davis, J. F. (1952). Manual of Surface Electromyography. Montreal: Laboratory for Psychological Studies, Allan Memorial Institute of Psychiatry.Google Scholar
Davis, J. F. & Malmo, R. B. (1951). Electromyographic recording during interview. American Journal of Psychiatry, 107: 908916.Google Scholar
Davis, J. F., Malmo, R. B., & Shagass, C. (1954). Electromyographic reaction to strong auditory stimulation in psychiatric patients. Canadian Journal of Psychology, 8: 177186.Google Scholar
Davis, M., Walker, D. L., & Myers, K. M. (2003). Role of the amygdala in fear extinction measured with potentiated startle. Annals of the New York Academy of Sciences, 985: 218232.Google Scholar
Davis, R. C. (1938). The relation of muscle action potentials to difficulty and frustration. Journal of Experimental Psychology, 23: 141158.Google Scholar
Davis, R. C. (1940). Set and Muscular Tension. Bloomington, IN: Indiana University Publications, Science series no. 10.Google Scholar
Davis, W. J., Rahman, M. A., Smith, L. J., Burns, A., Senecal, L., McArthur, D., … & Wagner, W. (1995). Properties of human affect induced by static color slides (IAPS): dimensional, categorical and electromyographic analysis. Biological Psychology, 41: 229253.Google Scholar
De Luca, C. J. & Erim, Z. (1994). Common drive of motor units in regulation of muscle force. Trends in Neurosciences, 17: 299305.Google Scholar
De Luca, C. J. & Knaflitz, M. (1992). Surface Electromyography: What’s New? Turin: C.L.U.T. Editrice.Google Scholar
Dennis, M., Agostino, A., Taylor, H. G., Bigler, E. D., Rubin, K., Vannatta, K., … & Yeates, K. O. (2013). Emotional expression and socially modulated emotive communication in children with traumatic brain injury. Journal of the International Neuropsychological Society, 19: 3443.Google Scholar
Desmedt, J. E. & Godaux, E. (1981). Spinal motoneuron recruitment in man: rank deordering with direction but not with speed of movement. Science, 214: 933936.Google Scholar
Detenber, B. H., Simmons, R. F., & Reiss, J. E. (2000). The emotional significance of color in television presentations. Media Psychology, 2: 331355.Google Scholar
Dimberg, U. (1982). Facial reactions to facial expressions. Psychophysiology, 19: 643647.Google Scholar
Dimberg, U. (1990). Gender differences in facial reactions to facial expressions. Biological Psychology, 30: 151159.Google Scholar
Disselhorst-Klug, C., Silny, J., & Rau, G. (1997). Improvement of spatial resolution in surface-EMG: a theoretical and experimental comparison of different spatial filters. IEEE Transactions on Biomedical Engineering, 44: 567574.Google Scholar
Dodge, R. B. (1911). A systematic exploration of a normal knee jerk, its technique, the form of muscle contraction, its amplitude, its latent time and its theory. Zeitschrift für Allgemeine Physiologie, 12: 158.Google Scholar
Dollins, A. B. & McGuigan, F. J. (1989). Frequency analysis of electromyographically measured covert speech behavior. Pavlovian Journal of Biological Science, 24: 2730.Google Scholar
Dorfman, D. & Cacioppo, J. T. (1990). Waveform moment analysis: topographical analysis of nonrhythmic waveforms. In Cacioppo, J. T. & Tassinary, L. G. (eds.), Principles of Psychophysiology (pp. 661707). Cambridge University Press.Google Scholar
Du, S., Tao, Y., & Martinez, A. (2014). Compound facial expressions of emotion. Proceedings of the National Academy of Sciences of the USA, 111: 14541462.Google Scholar
Duchenne, G. B. (1990 [1862]). The Mechanism of Human Facial Expression, ed. and trans. Cuthbertson, R. A.. Cambridge University Press.Google Scholar
Duffy, E. (1962). Activation and Behavior. New York: John Wiley.Google Scholar
Eason, R. G. & White, C. T. (1961). Muscular tension, effort, and tracking difficulty: studies of parameters which affect tension levels and performance efficiency. Perceptual and Motor Skills, 12: 331372.Google Scholar
Eccles, J. C. & Sherrington, C. S. (1930). Number and contraction values of individual motor-units examined in some muscles of the limb. Proceedings of the Royal Society of London, 106B: 326357.Google Scholar
Eibl-Eibesfeldt, I. (1972). Similarities and differences between cultures in expressive movement. In Hinde, R. A. (ed.), Nonverbal Communication (pp. 297314). Cambridge University Press.Google Scholar
Ekman, P. (1972). Universal and cultural differences in facial expressions of emotion. In Cole, J. (ed.), Nebraska Symposium on Motivation, vol. 19 (pp. 207218). Lincoln, NE: University of Nebraska Press.Google Scholar
Ekman, P., Friesen, W. V., & Hager, J. C. (2002). Facial Action Coding System [E-book]. Salt Lake City, UT: Research Nexus.Google Scholar
Englis, B. G., Vaughan, K. B., & Lanzetta, J. T. (1982). Conditioning of counter-empathic emotional responses. Journal of Experimental Social Psychology, 18: 375391.Google Scholar
Enoka, R. M. & Pearson, K. (2013). The motor unit and muscle action. In Kandel, E. R., Schwartz, J. H., Jessel, T. M., Sieglbaum, S. A., & Hudspeth, A. J. (eds.), Principles of Neural Science, 5th edn. (pp. 768789). New York: Elsevier.Google Scholar
Epstein, L. H. (1990). Perception of activity in the zygomaticus major and corrugator supercilii muscle regions. Psychophysiology, 27: 6872.Google Scholar
Erlacher, D. & Schredl, M. (2008). Do REM (lucid) dreamed and executed actions share the same neural substrate? International Journal of Dream Research, 1: 713.Google Scholar
Fanardjian, V. V. & Manvelyan, L. R. (1987). Mechanisms regulating the activity of facial nucleus motoneurons – IV. Influences from the brainstem structures. Neuroscience, 20: 845853.Google Scholar
Farina, D., Cescon, C., & Merletti, R. (2002). Influence of anatomical, physical, and detection-system parameters on surface EMG. Biological Cybernetics, 86: 445456.Google Scholar
Fearing, F. (1930). Reflex Action: A Study in the History of Physiological Psychology. Baltimore, MD: Williams & Wilkins.Google Scholar
Foerster, J. & Strack, F. (1997). Motor action in retrieval of valenced information: a motor congruence effect. Perceptual and Motor Skills, 85: 14191427.Google Scholar
Freeman, G. L. (1931). Mental activity and the muscular processes. Psychological Review, 38: 428447.Google Scholar
Fridlund, A. J. (1991). Sociality of solitary: potentiation by an implicit audience. Journal of Personality and Social Psychology, 60: 229240.Google Scholar
Fridlund, A. J. (1994). Human Facial Expression: An Evolutionary View. San Diego, CA: Academic Press.Google Scholar
Fridlund, A. J. & Cacioppo, J. T. (1986). Guidelines for human electromyographic research. Psychophysiology, 23: 567589.Google Scholar
Fridlund, A. J., Fowler, S. C., & Pritchard, D. A. (1980). Striate muscle tensional patterning in frontalis EMG biofeedback. Psychophysiology, 17: 4755.Google Scholar
Fridlund, A. J., Hatfield, M. E., Cottam, G. L., & Fowler, S. C. (1986). Anxiety and striate-muscle activation: evidence from electromyographic pattern analysis. Journal of Abnormal Psychology, 95: 228236.Google Scholar
Friesen, W. V. (1972). Cultural differences in facial expression in a social situation: an experimental text of the concept of display rules. Unpublished doctoral dissertation, University of California, San Francisco.Google Scholar
Fulton, J. F. (1926). Muscular Contraction and the Reflex Control of Movement. Baltimore, MD: Williams & Wilkins.Google Scholar
Gallese, V. (2003). The roots of empathy: the shared manifold hypothesis and the neural basis of intersubjectivity. Psychopathology, 36: 171180.Google Scholar
Gallistel, C. R. (1980). The Organization of Action: The New Synthesis. Hillsdale, NJ: Lawrence Erlbaum Associates.Google Scholar
Ganesh, G., Franklin, D. W., Gassert, R., Imamizu, H., & Kawato, M. (2007). Accurate real-time feedback of surface EMG during fMRI. Journal of Neurophysiology, 97: 912920.Google Scholar
Gans, C. & Gorniak, G. C. (1980). Electromyograms are repeatable: precautions and limitations. Science, 210: 795797.Google Scholar
Gazzoni, M. (2010). Multichannel surface electromyography in ergonomics: potentialities and limits. Human Factors and Ergonomics in Manufacturing & Service Industries, 20: 255271.Google Scholar
Geen, R. G. & Gange, J. J. (1977). Drive theory of social facilitation: twelve years of theory and research. Psychological Bulletin, 84: 12671288.Google Scholar
Geen, T. R. (1992). Facial expressions in socially isolated primates: open and closed programs for expressive behavior. Journal of Research in Personality, 26: 273280.Google Scholar
Geen, T. R. & Tassinary, L. G. (2002). The mechanization of expression in John Bulwar’s Pathomyotonia. American Journal of Psychology, 115: 275300.Google Scholar
Gehricke, J., & Shapiro, D. (2001). Facial and autonomic activity in depression: social context differences during imagery. International Journal of Psychophysiology, 41: 5364.Google Scholar
Geisser, M. E., Ranavaya, M., Haig, A. J., Roth, R. S., Zucker, R., Ambroz, C., & Caruso, M. (2005). A meta-analytic review of surface electromyography among persons with low back pain and normal, healthy controls. The Journal of Pain, 6: 711726.Google Scholar
Germana, J. (1974). Electromyography: human and general. In Thompson, R. F. & Patterson, M. M. (eds.), Bioelectric Recording Techniques: Part C: Receptor and Effector Processes (pp. 155163). New York: Academic Press.Google Scholar
Giggins, O. M., Persson, U. M., & Caulfield, B. (2013). Biofeedback in rehabilitation. Journal of Neuroengineering and Rehabilitation, 10(1): 6071.Google Scholar
Gilbert, A. N., Fridlund, A. J., & Sabini, J. (1987). Hedonic and social determinants of facial displays to odors. Chemical Senses, 12: 355363.Google Scholar
Gilbert, C. & Moss, D. P. (2003). Biofeedback and biological monitoring. In Moss, D. P., McGrady, A. V., Davies, T. C., & Wickremasekera, I. (eds.), Handbook of Mind-Body Medicine for Primary Care (pp. 109122). Thousand Oaks, CA: Sage.Google Scholar
Girard, E., Tassinary, L. G., Kappas, A., Gosselin, P., & Bontempo, D. (1997). The covert-to-overt threshold for facial actions: an EMG study. Psychophysiology, 34: S38.Google Scholar
Goldstein, J. B. (1972). Electromyography: a measure of skeletal muscle response. In Greenheld, N. S. & Sternbach, R. A. (eds.), Handbook of Psychophysiology (pp. 329366). New York: Holt, Rinehart & Winston.Google Scholar
Golla, F. & Antonovitch, S. (1929). The relaxation of muscular tonus and the patellar reflex to mental work. Journal of Mental Science, 75: 234241.Google Scholar
Gousain, A. K., Amarante, M. T. J., Hydem, J. S., & Yousif, N. J. (1996). A dynamic analysis of changes in the nasolabial fold using magnetic resonance imaging: implications for facial rejuvination and facial animation surgery. Plastic and Reconstructive Surgery, 98: 622636.Google Scholar
Graham, J. L. (1980). A new system for measuring nonverbal responses to marketing appeals. 1980 AMA Educator’s Conference Proceedings, 46: 340343.Google Scholar
Gray, H. (2000 [1918]). Anatomy of the Human Body, 20th edn. New York: Bartleby.com.Google Scholar
Groff, B. D., Baron, R. S., & Moore, D. L. (1983). Distraction, attentional conflict, and drivelike behavior. Journal of Experimental Social Psychology, 19: 359380.Google Scholar
Hackley, S. (1999). Implications of blink reflex research for theories of attention and consciousness. In Dawson, M. E., Schell, A. M., & Böhmelt, A. H. (eds.), Startle Modification: Implications for Neuroscience, Cognitive Science, and Clinical Science (pp. 137156). Cambridge University Press.Google Scholar
Hager, J. C. & Ekman, P. (1981). Methodological problems in Tourangeau and Ellsworth’s study of facial expression and experience of emotion. Journal of Personality and Social Psychology, 40: 358362.Google Scholar
Hamm, J., Kohler, C. G., Gur, R. C., & Verma, R. (2011). Automated facial action coding system for dynamic analysis of facial expressions in neuropsychiatric disorders. Journal of Neuroscience Methods, 200: 237256.Google Scholar
Harmon-Jones, E. & Winkielman, P. (2007). Social Neuroscience: Integrating Biological and Psychological Explanations of Social Behavior. New York: Guilford Press.Google Scholar
Harvey, R. & Peper, E. (1997). Surface electromyography and mouse use. Ergonomics, 40: 781789.Google Scholar
Hatfield, E., Cacioppo, J. T., & Rapson, R. L. (1993). Emotional Contagion. Cambridge University Press.Google Scholar
Hayes, K. J. (1960). Wave analyses of tissue noise and muscle action potential. Journal of Applied Physiology, 15: 749752.Google Scholar
Hazlett, R. L. & Hazlett, S. Y. (1999). Emotional response to television commercials: facial EMG vs. self-report. Journal of Advertising Research, 39: 723.Google Scholar
Hefferline, R. F., Keenan, B., & Harford, R. A. (1959). Escape and avoidance conditioning in human subjects without their observation of the response. Science, 130: 13381339.Google Scholar
Henneman, E. (1980). Organization of the motoneuron pool: the size principle. In Mountcastle, V. E. (ed.), Medical Physiology, Volume 1, 14th edn. (pp. 718741). St. Louis, MO: Mosby.Google Scholar
Hermens, H. J., Freriks, B., Disselhorst-Klug, C., & Rau, G. (2000). Development of recommendations for SEMG sensors and sensor placement procedures. Journal of Electromyography and Kinesiology, 10: 361374.Google Scholar
Hess, U., Banse, R., & Kappas, A. (1995). The intensity of facial expression is determined by underlying affective state and social situation. Journal of Personality and Social Psychology, 69: 280288.Google Scholar
Hess, U., Kappas, A., McHugo, G. J., & Kleck, R. E. (1989). An analysis of the encoding and decoding of spontaneous and posed smiles: the use of facial electromyography. Journal of Nonverbal Behavior, 13: 121137.Google Scholar
Hess, U., Philippot, P., & Blairy, S. (1999). Mimicry: facts and fiction. In Philippot, P., Feldman, R., & Coats, E. (eds.), The Social Context of Nonverbal Behavior: Studies in Emotion and Social Interaction (pp. 213241). Cambridge University Press.Google Scholar
Hill, A. V. (1959). The heat production of muscle and nerve, 1848–1914. Annual Review of Physiology, 21: 118.Google Scholar
Hislop, H. J. & Montgomery, J. (2002). Daniels and Worthingham’s Muscle Testing: Techniques of Manual Examination, 7th edn. Philadelphia: W. B. Saunders Company.Google Scholar
Honts, C. R., Devitt, M. K., Winbush, M., & Kircher, J. C. (1996). Mental and physical countermeasures reduce the accuracy of the concealed knowledge test. Psychophysiology, 33: 8492.Google Scholar
Honts, C. R., Hodes, R. L., & Raskin, D. C. (1985). Effects of physical countermeasures on the physiological detection of deception. Journal of Applied Psychology, 79: 177187.Google Scholar
Honts, C. R., Raskin, D. C., & Kircher, J. C. (1987). Effects of physical countermeasures and their electromyographic detection during polygraph tests for deception. Journal of Psychophysiology, 1: 241247.Google Scholar
Honts, C. R., Raskin, D. C., & Kircher, J. C. (1994). Mental and physical countermeasures reduce the accuracy of polygraph tests. Journal of Applied Psychology, 79: 252259.Google Scholar
Hu, C. S., Wang, Q., Han, T., Weare, E., & Fu, G. (2015). Differential emotion attribution to neutral faces of own and other races. Cognition & Emotion. doi: 10.1080/02699931.2015.1092419.Google Scholar
Hu, S. & Wan, H. (2003). Imagined events with specific emotional valence produce specific patterns of facial EMG activity. Perceptual and Motor Skills, 97: 10911099.Google Scholar
Huang, C. N., Chen, C. H., & Chung, H. Y. (2006). Application of facial electromyography in computer mouse access for people with disabilities. Disability & Rehabilitation, 28: 231237.Google Scholar
Humphrey, G. (1951). Thinking. New York: John Wiley.Google Scholar
Hutchinson, R. R., Pierce, G. E., Emley, G. S., Proni, T. J., & Sauer, R. A. (1977). The laboratory measurement of human anger. Biobehavioral Reviews, 1: 241259.Google Scholar
Huxley, A. F. (1980). Reflections on Muscle. Liverpool University Press.Google Scholar
Isley, C. L. & Basmajian, J. V. (1973). Electromyography of the human cheeks and lips. Anatomical Record, 176: 143148.Google Scholar
Ison, J. R. & Hoffman, H. S. (1983). Reflex modification in the domain of startle: II. The anomalous history of a robust and ubiquitous phenomenon. Psychological Bulletin, 94: 317.Google Scholar
Izard, C. E. (1971). The Face of Emotion. New York: Appleton-Century-Crofts.Google Scholar
Jacobsen, A., Kales, A., Lehmann, D., & Hoedmaker, F. S. (1964). Muscle tonus in human subjects during sleep and dreaming. Experimental Neurology, 10: 418424.Google Scholar
Jacobsen, A., Kales, A., Zweizig, J. R., & Kales, J. (1965). Special EEG and EMG techniques for sleep research. American Journal of EEG Technology, 18: 510.Google Scholar
Jacobson, E. (1927). Action currents from muscular contractions during conscious processes. Science, 66: 403.Google Scholar
Jacobson, E. (1932). Electrophysiology of mental activities. American Journal of Psychology, 44: 677694.Google Scholar
James, W. (1884). What is an emotion?Mind, 9: 188205.Google Scholar
James, W. (1890). The Principles of Psychology. New York: Holt.Google Scholar
Jäncke, L. (1996). Facial EMG in an anger-provoking situation: individual differences in directing anger outwards or inwards. International Journal of Psychophysiology, 23: 207214.Google Scholar
Jäncke, L. & Kaufmann, N. (1994). Facial EMG responses to odors in solitude and with an audience. Chemical Senses, 19: 99111.Google Scholar
Jenny, A. B. & Saper, C. B. (1987). Organization of the facial nucleus and corticofacial projection in the monkey: a reconsideration of the upper motor neuron facial palsy. Neurology, 37: 930939.Google Scholar
Jensen, R. (1999). Pathophysiological mechanisms of tension-type headache: a review of epidemiological and experimental studies. Cephalagia, 19: 602621.Google Scholar
Kendall, F. & McCreary, E. K. (1993). Muscles: Testing and Function, 4th edn. Baltimore, MD: Williams & Wilkins.Google Scholar
Khan, S. D., Bloodworth, D. S., & Woods, R. H. (1971). Comparative advantages of bipolar abraded skin surface electrodes over bipolar intramuscular electrodes for single motor unit recording in psychophysiological research. Psychophysiology, 8: 635647.Google Scholar
Kleinke, C. L. & Walton, J. H. (1982). Influence of reinforced smiling on affective responses in an interview. Journal of Personality and Social Psychology, 42: 557565.Google Scholar
Knowlton, B. J. & Squire, L. R. (1993). The learning of categories: parallel brain systems for item memory and category knowledge. Science, 262: 17471749.Google Scholar
Komi, P. V. & Buskirk, E. R. (1970). Reproducibility of electromyographic measurements with inserted wire electrodes and surface electrodes. Electromyography, 10: 357367.Google Scholar
Kraut, R. E. & Johnson, R. E. (1979). Social and emotional messages of smiling: an ethological approach. Journal of Personality and Social Psychology, 37: 15391553.Google Scholar
Kreibig, S. D., Samson, A. C., & Gross, J. J. (2013). The psychophysiology of mixed emotional states. Psychophysiology, 50: 799811.Google Scholar
Kritikos, A. & Brasch, C. (2008). Visual and tactile integration in action comprehension and execution. Brain Research, 1242: 7386.Google Scholar
Kuroda, Y., Thatcher, J., & Thatcher, R. (2011). Metamotivational state and dominance: links with EMG gradients during isokinetic leg extension and a test of the misfit effect. Journal of Sports Science, 29: 403410.Google Scholar
Laird, J. D. (1984). The real role of facial response in the experience of emotion: a reply to Tourangeau and Ellsworth, and others. Journal of Personality and Social Psychology, 47: 909917.Google Scholar
Lamotte, T., Priez, A., Lepoivre, E., Duchêne, J., & Tarrière, C. (1996). Surface electromyography as a tool to study head rest comfort in cars. Ergonomics, 39: 781796.Google Scholar
Landis, C. & Hunt, W. A. (1937). Magnification of time as a research technique in the study of behavior. Science, 85: 384385.Google Scholar
Lang, P. J., Greenwald, M. K., Bradley, M. M., & Hamm, A. O. (1993). Looking at pictures: affective, facial, visceral, and behavioral reactions. Psychophysiology, 30: 261273.Google Scholar
Lanzetta, J. T. & Englis, B. G. (1989). Expectations of cooperation and competition and their effects on observers’ vicarious emotional responses. Journal of Personality and Social Psychology, 56: 543554.Google Scholar
Lapatki, B. G., Oostenveld, R., Van Dijk, J. P., Jonas, I. E., Zwarts, M. J., and Stegeman, D. F. (2010). Optimal placement of bipolar surface EMG electrodes in the face based on single motor unit analysis, Psychophysiology, 47: 299314.Google Scholar
Larsen, J. T., Norris, C. J., & Cacioppo, J. T. (2003). Effects of positive and negative affect on electromyographic activity over zygomaticus major and corrugator supercilii. Psychophysiology, 40: 776785.Google Scholar
Larsen, R. J., Kasimatis, M., & Frey, K. (1992). Facilitating the furrowed brow: an unobtrusive test of the facial feedback hypothesis applied to unpleasant affect. Cognition & Emotion, 6: 321338.Google Scholar
Laurenti-Lions, L., Gallego, J., Chambille, B., Vardon, G., & Jacquemin, C. (1985). Control of myoelectrical responses through reinforcement. Journal of the Experimental Analysis of Behavior, 44: 185193.Google Scholar
Leifting, B., Bes, F., Fagioli, I., & Salzarulo, P. (1994). Electromyographic activity and sleep states in infants. Sleep, 17: 718722.Google Scholar
Leite, I., Pereira, A., Mascarenhas, S., Martinho, C., Prada, R., & Paiva, A. (2013). The influence of empathy in human–robot relations. International Journal of Human–Computer Studies, 71: 250260.Google Scholar
Levenson, R. W. (1992). Autonomic nervous system differences among emotions. Psychological Science, 3: 2327.Google Scholar
Levenson, R. W., Ekman, P., & Friesen, W. (1990). Voluntary facial action generates emotion-specific autonomic nervous system activity. Psychophysiology, 27: 363384.Google Scholar
Lew, H. L., Johnson, E. W., & Pease, W. S. (2005). Johnson’s Practical Electromyography, 4th edn. Baltimore, MD: Lippincott Williams & Wilkins.Google Scholar
Liddell, E. G. T. (1960). The Discovery of the Reflexes. Oxford: Clarendon Press.Google Scholar
Liddell, E. G. T. & Sherrington, C. S. (1925). Recruitment and some other features of reflex inhibition. Proceedings of the Royal Society of London (Biology), 97: 488518.Google Scholar
Lindsley, D. B. (1935). Electrical activity of human motor units during voluntary contraction. American Journal of Physiology, 114: 9099.Google Scholar
Lippold, O. C. J. (1967). Electromyography. In Venables, P. H. & Martin, I. (eds.), Manual of Psychophysiological Methods (pp. 245298). New York: John Wiley.Google Scholar
Littlewort, G. C., Bartlett, M. S., & Lee, K. (2009). Automatic coding of facial expressions displayed during posed and genuine pain. Image and Vision Computing, 27: 17971803.Google Scholar
Loeb, G. E. & Gans, C. (1986). Electromyography for Experimentalists. University of Chicago Press.Google Scholar
Lowery, M. M., Stoykov, N. S., & Kuiken, T. A. (2003). Independence of myoelectric control signals examined using a surface EMG model. IEEE Transactions on Biomedical Engineering, 50: 789793.Google Scholar
Lundberg, U., Dohns, I. E., Melin, B., Sandsjo, L., Palmerud, G., Kadefors, R., … & Parr, D. (1999). Psychophysiological stress responses, muscle tension, and neck and shoulder pain among supermarket cashiers. Journal of Occupational Health Psychology, 4: 245255.Google Scholar
Lundqvist, L. O. & Dimberg, U. (1995). Facial expressions are contagious. Journal of Psychophysiology, 9: 203211.Google Scholar
Luria, A. R. (1932). The Nature of Human Conflicts. New York: Liveright.Google Scholar
Lutz, R. S. (2003). Covert muscle excitation is outflow from the central generation of motor imagery. Behavioural Brain Research, 140: 149163.Google Scholar
Lykken, D. (1998). A Tremor in the Blood: Uses and Abuses of the Lie Detector, 2nd edn. New York: Plenum.Google Scholar
Lynn, P. A., Bettles, N. D., Hughes, A. D., & Johnson, S. W. (1978). Influences of electrode geometry on bipolar recordings of the surface electromyogram. Medical and Biological Engineering and Computing, 16: 651660.Google Scholar
Ma, C., Szeto, G. P., Yan, T., Wu, S., Lin, C., & Li, L. (2011). Comparing biofeedback with active exercise and passive treatment for the management of work-related neck and shoulder pain: a randomized controlled trial. Archives of Physical Medicine and Rehabilitation, 92: 849858.Google Scholar
Madsen, M., El Kaliouby, R., Eckhardt, M., Hoque, M. E., Goodwin, M. S., & Picard, R. (2009). Lessons from participatory design with adolescents on the autism spectrum. In CHI’09 Extended Abstracts on Human Factors in Computing Systems (pp. 38353840). New York: ACM.Google Scholar
Malcolm, R., Von, J. M., & Horney, R. A. (1989). Correlations between facial electromyography and depression. Psychiatric Forum, 15: 1923.Google Scholar
Malmo, R. B. (1965). Physiological gradients and behavior. Psychological Bulletin, 64: 225234.Google Scholar
Malmo, R. B. (1975). On Emotions, Needs, and Our Archaic Brain. New York: Holt, Rinehart & Winston.Google Scholar
Malmo, R. B. & Malmo, H. P. (2000). On electromyographic (EMG) gradients and movement-related brain activity: significance for motor control, cognitive functions, and certain psychopathologies. International Journal of Psychophysiology, 38: 145209.Google Scholar
Markovsky, B. & Berger, S. M. (1983). Crowd noise and mimicry. Personality and Social Psychology Bulletin, 9: 9096.Google Scholar
Martin, I. (1956). Levels of muscle activity in psychiatric patients. Acta Psychologica, 12: 326341.Google Scholar
Mathews, B. H. C. (1934). A special purpose amplifier. Journal of Physiology (London), 81: 28.Google Scholar
Max, L. W. (1932). Myoesthesis and “imageless thought.” Science, 76: 235236.Google Scholar
Max, L. W. (1937). An experimental study of the motor theory of consciousness: IV. Action-current responses in the deaf during awakening, kinaesthetic imagery and abstract thinking. Journal of Comparative Psychology, 24: 301344.Google Scholar
McCanne, T. R. & Anderson, J. A. (1987). Emotional responding following experimental manipulation of facial electromyographic activity. Journal of Personality and Social Psychology, 52: 759768.Google Scholar
McClelland, J. L. (1979). On the time relations of mental processes: an examination of systems in cascade. Psychological Review, 86: 287330.Google Scholar
McGarry, T. & Franks, I. (1997). A horse race between independent processes: evidence for a phantom point of no return in the preparation of a speeded motor response. Journal of Experimental Psychology: Human Perception and Performance, 23: 15331542.Google Scholar
McGuigan, F. J. (1966). Thinking: Studies of Covert Language Processes. New York: Appleton-Century-Crofts.Google Scholar
McGuigan, F. J. (1970). Covert oral behavior during the silent performance of language. Psychological Bulletin, 74: 309326.Google Scholar
McGuigan, F. J. (1978). Cognitive Psychophysiology: Principles of Covert Behavior. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
McGuigan, F. J. & Bailey, S. C. (1969). Logitudinal study of covert oral behavior during silent reading. Perceptual and Motor Skills, 28: 170.Google Scholar
McHugo, G. & Lanzetta, J. T. (1983). Methodological decisions in social psychophysiology. In Cacioppo, J. T. & Petty, R. E. (eds.), Social Psychophysiology: A Sourcebook (pp. 630665). New York: Guilford Press.Google Scholar
Merletti, R. & Parker, P. A. (2004). Electromyography: Physiology, Engineering, and Non-Invasive Applications. New York: John Wiley.Google Scholar
Merton, P. A. (2004). Reflexes. In Gregory, R. L. (ed.), The Oxford Companion to the Mind. Oxford University Press.Google Scholar
Mesin, L., Merletti, R., & Rainoldi, A. (2009). Surface EMG: the issue of electrode location. Journal of Electromyography and Kinesiology, 19: 719726.Google Scholar
Meyer, D. R. (1953). On the interaction of simultaneous responses. Psychological Bulletin, 20: 204220.Google Scholar
Meyer, D. R., Bahrick, H. P., & Fitts, P. M. (1953). Incentive, anxiety, and the human blink rate. Journal of Experimental Psychology, 45: 183287.Google Scholar
Morecraft, R. J., Stilwell-Morecraft, K. S., & Rossing, W. R. (2004). The motor cortex and facial expression: new insights from neuroscience. The Neurologist, 10: 235249.Google Scholar
Mulder, T. & Hulstijn, W. (1984). The effect of fatigue and repetition of the task on the surface electromyographic signal. Psychophysiology, 21: 528534.Google Scholar
Nakamura, T., Maejima, A., & Morishima, S. (2013). Detection of driver’s drowsy facial expression. In Proceedings of the 2nd IAPR Asian Conference on Pattern Recognition (ACPR), 2013 (pp. 749753). Piscataway: IEEE.Google Scholar
Needham, D. M. (1971). Machina Carnis: The Biochemistry of Muscular Contraction in its Historical Development. London: Cambridge University Press.Google Scholar
Nestoriuc, Y., Martin, A., Rief, W., & Andrasik, F. (2008). Biofeedback treatment for headache disorders: a comprehensive efficacy review. Applied Psychophysiology and Biofeedback, 33: 125140.Google Scholar
Nestoriuc, Y., Rief, W., & Martin, A. (2008). Meta-analysis of biofeedback for tension-type headache: efficacy, specificity, and treatment moderators. Journal of Consulting and Clinical Psychology, 76: 379396.Google Scholar
Netter, F. (1991). Anatomy, Physiology and Metabolic Disorders. Part 1: Musculoskeletal System. Ciba Collection, vol. 8. Chichester: John Wiley.Google Scholar
Neumann, D. L., Lipp, O. V., & Pretorius, N. R. (2004). The effects of lead stimulus and reflex stimulus modality on modulation of the blink reflex at very short, short, and long lead intervals. Perception and Psychophysics, 66: 141151.Google Scholar
Neumann, R. & Strack, F. (2000). Experiential and nonexperiential routes of motor influence on affect and evaluation. In Bless, H. & Forgas, J. P. (eds.), The Message Within: The Role of Subjective Experience in Social Cognition and Behavior (pp. 5268). Philadelphia, PA: Psychology Press.Google Scholar
Nosofsky, R. M. & Zaki, S. R. (1998). Dissociations between categorization and recognition in amnesic and normal individuals: an exemplar-based interpretation. Psychological Science, 9: 247255.Google Scholar
Oddsson, L. I. E., Giphart, J. E., Buijs, R. J. C., Roy, S. H., Taylor, H. P., & De Luca, C. J. (1997). Development of new protocols and analysis procedures for the assessment of LBP by surface EMG techniques. Journal of Rehabilitation Research and Development, 34: 415426.Google Scholar
O’Dwyer, N. J., Quinn, P. T., Guitar, B. E., Andrews, G., & Neilson, P. D. (1981). Procedures for verification of electrode placement in EMG studies of orofacial and mandibular muscles. Journal of Speech and Hearing Research, 241: 273288.Google Scholar
Olivers, C. N. L. & Nieuwenhuis, S. (2005). The beneficial effect of concurrent task-irrelevant mental activity on temporal attention. Psychological Science, 16: 265269.Google Scholar
Ota, N., Gahr, M., & Soma, M. (2015). Tap dancing birds: the multimodal mutual courtship display of males and females in a socially monogamous songbird. Scientific Reports, 5, Article 16614. Retrieved November 23, 2015 from http://dx.doi.org/10.1038/srep16614.Google Scholar
Paloheimo, M. (1990). Quantitative surface electromyography (qEMG): applications in anaesthesiology and critical care. Acta Anaesthesiologica Scandinavica [Supplementum 93], 34: 183.Google Scholar
Pernkopf, E. (1980). Atlas of Topographical and Applied Human Anatomy, 2nd edn., trans. Monsen, H., vol. 1. Philadelphia: W. B. Saunders Company.Google Scholar
Perry, T. J. & Goldwater, B. C. (1987). A passive behavioral measure of sleep onset in high-alpha and low alpha subjects. Psychophysiology, 24: 657665.Google Scholar
Petrides, M., Cadoret, G., & Mackey, S. (2005). Orofacial somatomotor responses in the macaque monkey homologue of Broca’s area. Nature, 435: 12351238.Google Scholar
Picard, R. W. (1997). Affective Computing. Cambridge, MA: MIT Press.Google Scholar
Pikoff, H. (1984). Is the muscular model of headache still viable?Headache, 24: 186198.Google Scholar
Pivik, R. T. (2007). Sleep and dreaming. In Cacioppo, J. T., Tassinary, L. G., & Berntson, G. (eds.), Handbook of Psychophysiology (pp. 633664). Cambridge University Press.Google Scholar
Platt, J. R. (1964). Strong inference. Science, 146: 347353.Google Scholar
Porter, S., ten Brinke, L., Baker, A., & Wallace, B. (2011). Would I lie to you? “Leakage” in deceptive facial expressions relates to psychopathy and emotional intelligence. Personality and Individual Difference, 51: 133137.Google Scholar
Porter, S., ten Brinke, L., & Wallace, B. (2012). Secrets and lies: involuntary leakage in deceptive facial expressions as a function of emotional intensity. Journal of Nonverbal Behavior, 36: 2337.Google Scholar
Pratt, F. H. (1917). The all-or-none principle in graded response of skeletal muscle. American Journal of Physiology, 44: 517542.Google Scholar
Pratt, F. H. & Eisenberger, J. P. (1919). The quantal phenomena in muscle: methods with further evidence of the all-or-none principle in graded response for the skeletal fibre. American Journal of Physiology, 49: 154.Google Scholar
Pressman, M. R. & Orr, W. C. (eds.) (1997). Understanding Sleep: The Evaluation and Treatment of Sleep Disorders. Washington, DC: American Psychological Association.Google Scholar
Pritchard, D. (1995). EMG levels in children who suffer from severe headache. Headache, 35: 554556.Google Scholar
Pullman, S. K., Goodin, D. S., Marquinez, A. I., Tabbal, S., & Rubin, M. (2000). Clinical utility of surface EMG: report of the therapeutics and technology assessment subcommittee of the American Academy of Neurology. Neurology, 55: 171177.Google Scholar
Rainoldi, A., Melchiorri, G., & Caruso, I. (2004). A method for positioning electrodes during surface EMG recordings in lower limb muscles. Journal of Neuroscience Methods, 134: 3743.Google Scholar
Rajecki, D. W. (1983). Animal aggression: implications for human aggression. In Geen, R. G. & Donnerstein, E. J. (eds.), Aggression: Theoretical and Empirical Reviews, Vol. 1 (pp. 189211). New York: Academic Press.Google Scholar
Rankin, R. E. & Campbell, D. (1955). Galvanic skin response to negro and white experimenters. Journal of Abnormal and Social Psychology, 51: 3033.Google Scholar
Ravaja, N., Kallinen, K., Saari, T., & Keltikangas-Jarvinen, L. (2004). Suboptimal exposure to facial expressions when viewing video messages from a small screen: effects on emotion, attention, and memory. Journal of Experimental Psychology: Applied, 10: 120131.Google Scholar
Rieger, R. & Deng, S. (2013). Double-differential recording and AGC using microcontrolled variable gain ASIC. IEEE Transactions on Neural Systems and Rehabilitation Engineering, 21: 4754.Google Scholar
Rimehaug, T. & Sveback, S. (1987). Psychogenic muscle tension: the significance of motivation and negative affect in perceptual-cognitive task performance. International Journal of Psychophysiology, 5: 97106.Google Scholar
Rissen, D., Melin, B., Sandsjo, L., Dohns, I., & Lundberg, U. (2000). Surface EMG and psychophysiological stress reactions in women during repetitive work. European Journal of Applied Physiology, 83: 215222.Google Scholar
Ritz, T., Dahme, B., & Claussen, C. (1999). Gradients of facial EMG and cardiac activity during emotional stimulation. Journal of Psychophysiology, 13: 317.Google Scholar
Rizzolatti, G., Fogassi, L., & Gallese, V. (2001). Neurophysiological mechanisms underlying the understanding and imitation of action. Nature Reviews Neuroscience, 2: 661670.Google Scholar
Rosenthal, R. & Rosnow, R. (eds.) (1969). Artifact in Behavioral Research. New York: Academic Press.Google Scholar
Rossi, A. M. (1959). An evaluation of the manifest anxiety scale by the use of electromyography. Journal of Experimental Psychology, 58: 6469.Google Scholar
Roy, S. H., De Luca, C., Emley, M., Oddsson, L. I. E., Buijis, R. J. C., Levins, J., … & Jabre, J. F. (1997). Classification of back muscle impairment based on the surface electromyographic signal. Journal of Rehabilitation Research and Development, 34: 405414.Google Scholar
Russell, J. A. & Fernández-Dols, J. M. (eds.) (1997). The Psychology of Facial Expression. Cambridge University Press.Google Scholar
Sacco, I. C. N., Gomes, A. A., Otuzi, M. E., Pripas, D., & Onodera, A. N. (2009). A method for better positioning bipolar electrodes for lower limb EMG recordings during dynamic contractions. Journal of Neuroscience Methods, 180: 133137.Google Scholar
Sato, W., Fujimura, T., Kochiyama, T., & Suzuki, N. (2013). Relationships among facial mimicry, emotional experience, and emotion recognition. PLoS One, 8: e57889. doi: 10.1371/journal.pone.0057889.Google Scholar
Schmidt-Nielsen, K. (1997). Animal Physiology: Adaptation and Environment, 5th edn. Cambridge University Press.Google Scholar
Schwartz, G. E., Fair, P. L., Salt, P., Mandel, M. R., & Klerman, G. L. (1976). Facial muscle patterning to affective imagery in depressed and nondepressed subjects. Science, 192: 489491.Google Scholar
Seiler, R. (1973). On the function of facial muscles in different behavioral situations: a study based on the muscle morphology and electromyography. American Journal of Physical Anthropology, 38: 567572.Google Scholar
Shaw, W. A. (1940). The relation of muscular action potentials to imaginal weight lifting. Archives of Psychology, 35: 150.Google Scholar
Shergill, G. S., Sarrafzadeh, A., Diegel, O., & Shekar, A. (2008). Computerized sales assistants: the application of computer technology to measure consumer interest – a conceptual framework. Journal of Electronic Commerce Research, 9: 176191.Google Scholar
Sherrington, C. S. (1923 [1906]). The Integrative Actions of the Nervous System. New Haven, CT: Yale University Press.Google Scholar
Shimizu, A. & Inoue, T. (1986). Dreamed speech and speech muscle activity. Psychophysiology, 23: 210215.Google Scholar
Sims, T. B., van Reekum, C. M., Johnstone, T., & Chakrabarti, B. (2012). How reward modulates mimicry: EMG evidence of greater facial mimicry of more rewarding faces. Psychophysiology, 49: 9981004.Google Scholar
Skinner, B. F. (1931). The concept of the reflex in the description of behavior. Journal of General Psychology: Experimental, Theoretical, Clinicial, and Historical Psychology, 5: 427457.Google Scholar
Sloan, D. M., Bradley, M. M., Dimoulas, E., & Lang, P. J. (2002). Looking at facial expressions: dysphoria and facial EMG. Biological Psychology, 60: 7990.Google Scholar
Smith, J. D. & Minda, J. P. (2001). Journey to the center of the category: the dissociation in amnesia between categorization and recognition. Journal of Experimental Psychology: Human Learning, Memory & Cognition, 27: 9841002.Google Scholar
Smith, R. R. & Kier, W. M. (1989). Trunks, tongues, and tentacles: moving with skeletons of muscle. American Scientist, 77: 2835.Google Scholar
Sonnby-Borgström, M. & Jönsson, P. (2003). Models-of-self and models-of-others as related to facial muscle reactions at different levels of cognitive control. Scandinavian Journal of Psychology, 44: 141151.Google Scholar
Spencer, H. (1870). Principles of Psychology, 2nd edn. London: Williams & Norgate.Google Scholar
Sperry, R. (1952). Neurology and the mind–brain problem. American Scientist, 40: 291312.Google Scholar
Stark, R., Walter, B., Schienle, A., & Vaitl, D. (2005). Psychophysiological correlates of disgust and disgust sensitivity. Journal of Psychophysiology, 19: 5060.Google Scholar
Sternbach, R. A. (1966). Principles of Psychophysiology. New York: Academic Press.Google Scholar
Sternberg, S. (1969). The discovery of processing stages: extensions of Donder’s method. Acta Psychologica, 30: 276315.Google Scholar
Stone, E. & Skubic, M. (2011). Evaluation of an inexpensive depth camera for in-home gait assessment. Journal of Ambient Intelligence and Smart Environments, 3: 349361.Google Scholar
Stoyva, J. & Budzynski, T. (1974). Cultivated low arousal: an anti-stress response? In DiCara, L. V. (ed.), Recent Advances in Limbic and Autonomic Nervous System Research (pp. 370394). New York: Plenum Press.Google Scholar
Strack, F., Martin, L. L., & Stepper, J. (1988). Inhibitory and facilitatory conditions of the human smile: a nonobtrusive test of the facial feedback hypothesis. Journal of Personality and Social Psychology, 54: 768777.Google Scholar
Stringham, J. M., Fuld, K., & Wenzel, A. J. (2003). Action spectrum for photophobia. Journal of the Optical Society of America, 20: 18521858.Google Scholar
Sumitsuji, N., Nan’no, H., Kuwata, Y., & Ohta, Y. (1980). The effects of the noise due to the jet airplane to the human facial expression (EMG study), EEG changes and their manual responses to the various sleeping stages of the subjects. Electromyography and Clinical Neurophysiology, 20: 4972.Google Scholar
Svebak, S., Dalen, K., & Storfjell, O. (1981). The psychological significance of task-induced tonic changes in somatic and autonomic activity. Psychophysiology, 18: 403409.Google Scholar
Swazey, J. P. (1969). Reflexes and Motor Integration: Sherrington’s Concept of Integrative Action. Cambridge, MA: Harvard University Press.Google Scholar
Tassinary, L. G. & Cacioppo, J. T. (1992). Unobservable facial actions and emotion. Psychological Science, 3: 2833.Google Scholar
Tassinary, L. G., Cacioppo, J. T., & Geen, T. R. (1989). A psychometric study of surface electrode placements for facial electromyographic recording: I. The brow and cheek muscle regions. Psychophysiology, 26: 116.Google Scholar
Tassinary, L. G., Cacioppo, J. T., & Vanman, E. (2007). The skeletomotor system: surface electromyography. In Cacioppo, J. T., Tassinary, L. G., and Bernston, G. (eds.), Handbook of Psychophysiology, 3rd edn. (pp. 267299). Cambridge University Press.Google Scholar
Tassinary, L. G., Orr, S. P., Wolford, G., Napps, S. E., & Lanzetta, J. T. (1984). The role of awareness in affective information processing: an exploration of the Zajonc hypothesis. Bulletin of the Psychonomic Society, 22: 489492.Google Scholar
Tassinary, L. G., Vanman, E., Geen, T. R., & Cacioppo, J. T. (1987). Optimizing surface electrode placements for facial EMG recordings: guidelines for recording from the perioral muscle region. Psychophysiology, 24: 615616.Google Scholar
ten Brinke, L., Porter, S., & Baker, A. (2012). Darwin the detective: observable facial muscles reveal emotional high-stakes lies. Evolution and Human Behavior, 33: 411416.Google Scholar
Teuber, H. L. (1955). Physiological psychology. Annual Review of Psychology, 6: 267294.Google Scholar
Thorson, A. M. (1925). The relation of tongue movements to internal speech. Journal of Experimental Psychology, 8: 132.Google Scholar
Tomovic, R. & Bellman, R. (1970). A systems approach to muscle control. Mathematical Biosciences, 8: 265277.Google Scholar
Tourangeau, R. & Ellsworth, P. C. (1979). The role of facial response in the experience of emotion. Journal of Personality and Social Psychology, 37: 15191531.Google Scholar
Trepman, E., Gellman, R. E., Solomon, R., Murthy, K. R., Micheli, L. J., & De Luca, C. (1994). Electromyographic analysis of standing posture and demi-plié in ballet and modern dancers. Medicine and Science in Sports and Exercise, 26: 771782.Google Scholar
Triplett, N. (1898). The dynamogenic factors in pacemaking and competition. American Journal of Psychology, 9: 507533.Google Scholar
Tyron, W. W. (1991). Activity Measurement in Psychology and Medicine. New York: Plenum Press.Google Scholar
van Boxtel, A. (2001). Optimal signal bandwidth for the recording of surface EMG activity of facial, jaw, oral, and neck muscles. Psychophysiology, 38: 2234.Google Scholar
van Boxtel, A. (2010). Facial EMG as a tool for inferring affective states. In Spink, A. J., Grieco, F., Krips, O. E., Loijens, L. W. S., Noldus, L. P. J. J., & Zimmerman, P. H. (eds.), Proceedings of Measuring Behavior 2010 (pp. 104108). Wageningen: Noldus Information Technology.Google Scholar
van Boxtel, A., Damen, E. J. P., & Brunia, C. H. M. (1996). Anticipatory EMG responses of the pericranial muscles in relation to heart rate during a warned simple reaction time task. Psychophysiology, 33: 576583.Google Scholar
van Boxtel, A., Goudswaard, P., & Janssen, K. (1983). Changes in EMG power spectra of facial and jaw-elevator muscles during fatigue. Journal of Applied Physiology, 54: 5158.Google Scholar
van Boxtel, A., Goudswaard, P., & Shomaker, L. R. B. (1984). Amplitude and bandwidth of the frontalis surface EMG: effects of electrode parameters. Psychophysiology, 21: 699707.Google Scholar
van Boxtel, A. & Jessurun, M. (1993). Amplitude and bilateral coherency of facial and jaw-elevator EMG activity as an index of effort during a two-choice serial reaction task. Psychophysiology, 30: 589604.Google Scholar
Van der Linden, J., Schoonderwaldt, E., Bird, J., & Johnson, R. (2011). Musicjacket: combining motion capture and vibrotactile feedback to teach violin bowing. IEEE Transactions on Instrumentation and Measurement, 60: 104113.Google Scholar
Vanman, E. J., Boehmelt, A. H., Dawson, M. E., & Schell, A. M. (1996). The varying time course of attentional and affective modulation of the startle eyeblink response. Psychophysiology, 33: 691697.Google Scholar
Vanman, E. J., Paul, B. Y., Ito, T. A., & Miller, N. (1997). The modern face of prejudice and structural features that moderate the effect of cooperation on affect. Journal of Personality and Social Psychology, 73: 941959.Google Scholar
Vanman, E. J., Saltz, J. L., Nathan, L. R., & Warren, J. A. (2004). Racial discrimination by low-prejudiced whites: facial movements as implicit measures of attitudes related to behavior. Psychological Science, 15: 711714.Google Scholar
Vaughan, K. B. & Lanzetta, J. T. (1980). Vicarious instigation and conditioning of facial expressive and autonomic responses to a model’s expressive display of pain. Journal of Personality and Social Psychology, 13: 909923.Google Scholar
Vaughan, K. B. & Lanzetta, J. T. (1981). The effects of modification of expressive displays on vicarious emotional arousal. Journal of Experimental Social Psychology, 17: 1630.Google Scholar
Vitti, M., Basmajian, J. V., Ouelette, P. L., Mitchell, D. L., Eastman, W. P., & Seaborn, R. D. (1975). Electromyographic investigations of the tongue and circumoral muscular sling with fine-wire electrodes. Journal of Dental Research, 54: 844849.Google Scholar
Vrana, S. R. & Rollock, D. (1998). Physiological response to a minimal social encounter: effects of gender, ethnicity, and social context. Psychophysiology, 35: 462469.Google Scholar
Vural, E., Bartlett, M., Littlewort, G., Cetin, M., Ercil, A., & Movellan, J. (2010). Discrimination of moderate and acute drowsiness based on spontaneous facial expressions. In Proceedings of the 2010 20th International Conference on Pattern Recognition (ICPR) (pp. 38743877). Piscataway, NJ: IEEE.Google Scholar
Waersted, M. & Westgaard, R. H. (1996). Attention-related muscle activity in different body regions during VDU work with minimal physical activity. Ergonomics, 39: 661676.Google Scholar
Warren, W. H. (2006). Dynamics of perception and action. Psychological Review, 113: 358389.Google Scholar
Wartenberg, R. (1946). The Examination of Reflexes: A Simplification. Chicago: Year Book Publishers.Google Scholar
Washburn, M. F. (1916). Movement and Imagery: Outlines of a Motor Theory of the Complexer Mental Processes. Boston, MA: Houghton Mifflin.Google Scholar
Waters, K. (1992). A physical model of facial tissue and muscle articulation derived from computer tomography data. Visualization in Biomedical Computing, 1808: 574583.Google Scholar
Weaver, C. V. (1977). Descriptive anatomical and quantitative variation in human facial musculature and the analysis of bilateral asymmetry. Dissertation Abstracts International (University Microfilms, no. 77-24, 305).Google Scholar
Wells, G. L. & Petty, R. E. (1980). The effects of overt head-movements on persuasion: compatibility and incompatibility of responses. Basic and Applied Social Psychology, 1: 219230.Google Scholar
Wexler, B. E., Warrenburg, S., Schwartz, G. E., & Jamner, L. D. (1992). EEG and EMG responses to emotion-evoking stimuli processed without conscious awareness. Neuropsychologia, 30: 10651079.Google Scholar
Wheatley, J. R., Tangel, D. J., Mezzanotte, W. S., & White, D. P. (1993). Influence of sleep on the alae nasi EMG and nasal resistance in normal man. Journal of Applied Physiology, 75: 626632.Google Scholar
Whitehill, J., Serpell, Z., Lin, Y., Foster, A., & Movellan, J. R. (2014). The faces of engagement: automatic recognition of student engagement from facial expressions. IEEE Transactions on Affective Computing, 5: 8698.Google Scholar
Wolf, K., Mass, R., Kiefer, F., Eckert, K., Weinhold, N., Wiedemann, K., & Naber, D. (2004). The influence of olanzapine on facial expressions of emotions in schizophrenia: an improved facial EMG study. German Journal of Psychiatry, 7: 1419.Google Scholar
Wolpert, D. (2011). The real reason for brains [Video file]. Retrieved from www.ted.com/talks/daniel_wolpert_the_real_reason_for_brains#t-69912Google Scholar
Wolpert, D., Pearson, K. G., & Ghez, C. (2013). The organization and planning of movement. In Kandel, E. R., Schwartz, J. H., Jessel, T. M., Sieglbaum, S. A., & Hudspeth, A. J. (eds.), Principles of Neural Science, 5th edn. (pp. 743767). New York: Elsevier.Google Scholar
Woodworth, R. S. & Schlosberg, H. (1954). Experimental Psychology, rev. edn. New York: Holt.Google Scholar
Wu, C. H. (1984). Electric fish and the discovery of animal electricity. American Scientist, 72: 598607.Google Scholar
Yucha, C. B. & Montgomery, D. (2008). Evidence-Based Practice in Biofeedback and Neurofeedback. Wheat Ridge, CO: Association for Applied Psychophysiology and Biofeedback.Google Scholar
Zajonc, R. B. (1965). Social facilitation. Science, 149: 269274.Google Scholar
Zipp, P. (1982). Recommendations for the standardization of lead positions in surface electromyography. European Journal of Applied Physiology, 50: 4154.Google Scholar

References

Allen, J. J., Chambers, A. S., & Towers, D. N. (2007). The many metrics of cardiac chronotropy: a pragmatic primer and a brief comparison of metrics. Biological Psychology, 74: 243262.Google Scholar
Ameloot, K., Palmers, P. J., & Malbrain, M. L. (2015). The accuracy of noninvasive cardiac output and pressure measurements with finger cuff: a concise review. Current Opinion on Critical Care, 21: 232239.Google Scholar
Anderson, C. R. (1998). Identification of cardiovascular pathways in the sympathetic nervous system. Clinical and Experimental Pharmacology and Physiology, 25: 449452.Google Scholar
Andersson, U. & Tracey, K. J. (2012). Reflex principles of immunological homeostasis. Annual Review of Immunology, 30: 313335.Google Scholar
Annila, P. A., Yli-Hankala, A. M., & Lindgren, L. (1994). The effect of atropine on the T-wave amplitude of ECG during isoflurane anaesthesia. International Journal of Clinical Monitoring and Computing, 11: 4347.Google Scholar
Armour, J. A. (2008). Potential clinical relevance of the “little brain” on the mammalian heart. Experimental Physiology, 93: 165176.Google Scholar
Babbs, C. F. (2012). Oscillometric measurement of systolic and diastolic blood pressures validated in a physiologic mathematical model. Biomedical Engineering Online, 11: 56.Google Scholar
Bachen, E. A., Manuck, S. B., Cohen, S., Muldoon, M. F., Raibel, R., Herbert, T. B., & Rabin, B. S. (1995). Adrenergic blockade ameliorates cellular immune responses to mental stress in humans. Psychosomatic Medicine, 57: 366372.Google Scholar
Bailey, R. H. & Bauer, J. H. (1993). A review of common errors in the indirect measurement of blood pressure. Archives of Internal Medicine, 153: 27412748.Google Scholar
Barbato, E. (2009). Role of adrenergic receptors in human coronary vasomotion. Heart, 95: 603608.Google Scholar
Barde, P. B., Jindal, G. D., Singh, R., & Deepak, K. K. (2006). New method of electrode placement for determination of cardiac output using impedance cardiography. Indian Journal of Physiology and Pharmacology, 50: 234240.Google 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: 4456.Google Scholar
Beaudin, A. E., Brugniaux, J. V., Vöhringer, M., Flewitt, J., Green, J. D., Friedrich, M. G., & Poulin, M. J. (2011). Cerebral and myocardial blood flow responses to hypercapnia and hypoxia in humans. American Journal of Physiology: Heart and Circulatory Physiology, 301: H1678H1686.Google Scholar
Beker, F., Weber, M., Fink, R. H., & Adams, D. J. (2003). Muscarinic and nicotinic ACh receptor activation differentially mobilize Ca2+ in rat intracardiac ganglion neurons. Journal of Neurophysiology, 90: 19561964.Google Scholar
Benschop, R. J., Nieuwenhuis, E. E. S., Tromp, E. A. M., Godart, G. L. R., Ballieux, R. E., & van Doornen, L. P. J. (1994). Effects of βadrenergic blockade on immunologic and cardiovascular changes induced by mental stress. Circulation, 89: 762769.Google Scholar
Bernstein, D. P. (1986). A new stroke volume equation for thoracic electrical bioimpedance: theory and rationale. Critical Care Medicine, 14: 904909.Google Scholar
Bernstein, D. P., Henry, I. C., Lemmens, H. J., Chaltas, J. L., DeMaria, A. N., Moon, J. B., & Kahn, A. M. (2015). Validation of stroke volume and cardiac output by electrical interrogation of the brachial artery in normals: assessment of strengths, limitations, and sources of error. Journal of Clinical Monitoring and Computing, 29: 789800.Google Scholar
Bernstein, D. P. & Lemmens, H. J. (2005). Stroke volume equation for impedance cardiography. Medical and Biological Engineering and Computers, 43: 443450.Google Scholar
Berntson, G. G., Bechara, A., Damasio, H., Tranel, D., Norman, G. J., & Cacioppo, J. T. (2011). The insula and evaluative processes. Psychological Science, 22: 8086.Google Scholar
Berntson, G. G., Bigger, J. T., Eckberg, D. L., Grossman, P., Kaufmann, P. G., Malik, M., … & van der Molen, M. W. (1997). Heart rate variability: origins, methods, and interpretive caveats. Psychophysiology, 34: 623648.Google Scholar
Berntson, G. G. & Cacioppo, J. T. (1999). Heart rate variability: a neuroscientific perspective for furthur studies. Cardiac Electrophysiology Review, 3: 279282.Google Scholar
Berntson, G. G. & Cacioppo, J. T. (2007). Integrative physiology: homeostasis, allostasis and the orchestration of systemic physiology. In Cacioppo, J. T., Tassinary, L. G., & Berntson, G. G. (eds.), Handbook of Psychophysiology, 3rd edn. (pp. 433452). Cambridge University Press.Google 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: 599608.Google Scholar
Berntson, G. G., Cacioppo, J. T., & Quigley, K. S. (1991). Autonomic determinism: the modes of autonomic control, the doctrine of autonomic space, and the laws of autonomic constraint. Psychological Review, 98: 459487.Google Scholar
Berntson, G. G., Cacioppo, J. T., & Quigley, K. S. (1993a). Cardiac psychophysiology and autonomic space in humans: empirical perspectives and conceptual implications. Psychological Bulletin, 114: 296322.Google Scholar
Berntson, G. G., Cacioppo, J. T., & Quigley, K. S. (1993b). Respiratory sinus arrhythmia: autonomic origins, physiological mechanisms, and psychophysiological implications. Psychophysiology, 30: 183196.Google Scholar
Berntson, G. G., Cacioppo, J. T., & Quigley, K. S. (1995). The metrics of cardiac chronotropism: biometric perspectives. Psychophysiology, 32: 162171.Google Scholar
Berntson, G. G., Lozano, D. L., & Chen, Y. J. (2005). Filter properties of the root mean square successive difference (RMSSD) statistic in heart rate. Psychophysiology, 42: 246252.Google Scholar
Berntson, G. G., Lozano, D. L., Chen, Y. J., & Cacioppo, J. T. (2004). Where to Q in PEP: reliability and validity. Psychophysiology, 41: 333337.Google Scholar
Berntson, G. G., Norman, G. J., Hawkley, L. C., & Cacioppo, J. T. (2008). Cardiac autonomic balance versus cardiac regulatory capacity. Psychophysiology, 45: 643652.Google Scholar
Bertinieri, G., di Rienzo, M., Cavallazzi, A., Ferrari, A. U., Pedotti, A., & Mancia, G. (1985). A new approach to analysis of the arterial baroreflex. Journal of Hypertension, 3: S79S81.Google Scholar
Billman, G. E. (2013). The LF/HF ratio does not accurately measure cardiac sympatho-vagal balance. Frontiers in Physiology, 4: article 26.Google Scholar
Borow, K. M. & Newberger, J. W. (1982). Noninvasive estimation of central aortic pressure using the oscillometric method for analyzing systemic artery pulsatile blood flow: comparative study of indirect systolic, diastolic and mean brachial artery pressure with simultaneous direct ascending aortic pressure measurements. American Heart Journal, 103: 879886.Google Scholar
Bosch, J. A., Berntson, G. G., Cacioppo, J. T., Dhabhar, F. S., & Marucha, P. T. (2003). Acute stress evokes a selective mobilization of T cells that differ in chemokine receptor expression: a potential pathway linking immunologic reactivity to cardiovascular disease. Brain, Behavior, & Immunity, 17: 251259.Google Scholar
Bosch, J. A., de Geus, E. J., Kelder, A., Veerman, E. C., Hoogstraten, J., & Amerongen, A. V. (2001). Differential effects of active versus passive coping on secretory immunity. Psychophysiology, 38: 836846.Google Scholar
Brack, K. E. (2015). The heart’s “little brain” controlling cardiac function in the rabbit. Experimental Physiology, 100: 348353.Google Scholar
Bresler, M. A., Sheffy, K., Pillar, G., Preiszler, M., & Herscovici, S. (2008). Differentiating between light and deep sleep stages using an ambulatory device based on peripheral arterial tonometry. Physiological Measurement, 29: 571584.Google Scholar
Brownley, K. A., Hurwitz, B. E., & Schneiderman, N. (2000). Cardiovascular psychophysiology. In Cacioppo, J. T., Tassinary, L. G., & Berntson, G. G. (eds.), Handbook of Psychophysiology, 2nd edn. (pp. 224264). Cambridge University Press.Google Scholar
Cacioppo, J. T. (1994). Social neuroscience: autonomic, neuroendocrine, and immune responses to stress. Psychophysiology, 31: 113128.Google Scholar
Cacioppo, J. T., Berntson, G. G., Binkley, P. F., Quigley, K. S., Uchino, B. N., & Fieldstone, A. (1994). Autonomic cardiac control: II. Basal response, noninvasive indices, and autonomic space as revealed by autonomic blockades. Psychophysiology, 31: 586598.Google Scholar
Cacioppo, J. T. & Hawkley, L. C. (2003). Social isolation and health, with an emphasis on underlying mechanisms. Perspectives in Biology and Medicine, 46 S39S52.Google Scholar
Cacioppo, J. T., Hawkley, L. C., Crawford, L. E., Ernst, J. M., Burleson, M. H., Kowalski, R. B., Malarkey, W. B., Van Cauter, E., & Berntson, G. G. (2002). Loneliness and health: potential mechanisms. Psychosomatic Medicine, 64: 407417.Google Scholar
Cacioppo, J. T., Malarkey, W. B., Kiecolt-Glaser, J. K., Uchino, B. N., Sgoutas-Emch, S. A., Sheridan, J. F., … & Glaser, R. (1995). Heterogeneity in neuroendocrine and immune responses to brief psychological stressors as a function of autonomic cardiac activation. Psychosomatic Medicine, 57: 154164.Google Scholar
Cacioppo, J. T., Tassinary, L. G., & Berntson, G. G. (eds.) (2007). Handbook of Psychophysiology. Cambridge University Press.Google Scholar
Capuana, L. J., Dywan, J., Tays, W. J., Elmers, J. L., Witherspoon, R., & Segalowitz, S. J. (2014). Factors influencing the role of cardiac autonomic regulation in the service of cognitive control. Biological Psychology, 102: 8889.Google Scholar
Cechetto, D. F. (2014). Cortical control of the autonomic nervous system. Experimental Physiology, 99: 326331.Google Scholar
Chin, K. Y. & Panerai, R. B. (2012). Comparative study of Finapres devices. Blood Pressure Monitoring, 17: 171178.Google Scholar
Chowdhary, S., Marsh, A. M., Coote, J. H., & Townend, J. N. (2004). Nitric oxide and cardiac muscarinic control in humans. Hypertension, 43: 10231028.Google Scholar
Chung, J. (2009). Echocardiography in 2009: state of the art. Journal of Invasive Cardiology, 21: 346351.Google Scholar
Cnockaert, L., Migeotte, P. F., Daubigny, L., Prisk, G. K., Grenez, F., & , R. C. (2008). A method for the analysis of respiratory sinus arrhythmia using continuous wavelet transforms. IEEE Transactions in Biomedical Engineering, 55: 16401642.Google Scholar
Cole, S. W., Hawkley, L. C., Arevalo, J. M., & Cacioppo, J. T. (2011). Transcript origin analysis identifies antigen-presenting cells as primary targets of socially regulated gene expression in leukocytes. Proceedings of the National Academy of Sciences of the USA, 108: 30803085.Google Scholar
Contrada, R. J. (1992). T-wave amplitude: on the meaning of a psychophysiological index. Biological Psychology, 33: 249258.Google Scholar
Corretti, M. C., Anderson, T. J., Benjamin, E. J., Celermajer, D., Charbonneau, F., Creager, M. A., … & Vogel, R. (2002). Guidelines for the ultrasound assessment of endothelial-dependent flow-mediated vasodilation of the brachial artery. Journal of the American College of Cardiology, 39: 257265.Google Scholar
Cotter, G., Schachner, A., Sasson, L., Dekel, H., & Moshkovitz, Y. (2006). Impedance cardiography revisited. Physiological Measures, 27: 817827.Google Scholar
Critchley, H. D., Nagai, Y., Gray, M. A., & Mathias, C. J. (2011). Dissecting axes of autonomic control in humans: insights from neuroimaging. Autonomic Neuroscience, 161: 3442.Google Scholar
Critchley, H. D., Rotshtein, P., Nagai, Y., O’Doherty, J., Mathias, C. J., & Dolan, R. J. (2005a). Activity in the human brain predicting differential heart rate responses to emotional facial expressions. NeuroImage, 24: 751762.Google Scholar
Critchley, H. D., Taggart, P., Sutton, P. M., Holdright, D. R., Batchvarov, V., Hnatkova, K., … & Dolan, R. J. (2005b). Mental stress and sudden cardiac death: asymmetric midbrain activity as a linking mechanism. Brain, 128: 7585.Google Scholar
Cybulski, G. (2011). Ambulatory Impedance Cardiography: The Systems and their Applications (Lecture Notes in Electrical Engineering). Berlin: Springer-Verlag.Google Scholar
Dampney, R. A., Polson, J. W., Potts, P. D., Hirooka, Y., & Horiuchi, J. (2003). Functional organization of brain pathways subserving the baroreceptor reflex: studies in conscious animals using immediate early gene expression. Cellular and Molecular Neurobiology, 23: 597616.Google Scholar
Davies, J. I. & Struthers, A. D. (2003). Pulse wave analysis and pulse wave velocity: a critical review of their strengths and weaknesses. Journal of Hypertension, 21: 463472.Google Scholar
De Vito, P. (2014). Atrial natriuretic peptide: an old hormone or a new cytokine? Peptides, 58: 108116.Google Scholar
deBoer, R. W., Karemaker, J. M., & Strackee, J. (1987). Hemodynamic fluctuations and baroreflex sensitivity in humans: a beat-to-beat model. American Journal of Physiology, 253: 680689.Google Scholar
Demeter, R. J., Parr, K. L., Toth, P. D., & Woods, J. R. (1993). Use of noninvasive bioelectric impedance to predict cardiac output in open heart recovery. Biological Psychology, 36: 2332.Google Scholar
Dessy, C., Moniotte, S., Ghisdal, P., Havaux, X., Noirhomme, P., & Balligand, J. L. (2004). Endothelial β3-adrenoceptors mediate vasorelaxation of human coronary microarteries through nitric oxide and endothelium-dependent hyperpolarization. Circulation, 110: 948954.Google Scholar
Di Rienzo, M., Parati, G., Castiglioni, P., Tordi, R., Mancia, G., & Pedotti, A. (2001). Baroreflex effectiveness index: an additional measure of baroreflex control of heart rate in daily life. American Journal of Physiology, 280: R744R751.Google Scholar
Docherty, J. R. (2002). Age-related changes in adrenergic neuroeffector transmission. Autonomic Neuroscience, 96: 812.Google Scholar
Eckberg, D. L. (1997). Sympathovagal balance: a critical appraisal. Circulation, 96: 32243232.Google Scholar
Eckberg, D. L. (1998). Sympathovagal balance: a critical appraisal – reply. Circulation, 98: 26432644.Google Scholar
Eckberg, D. L. (2000). Physiological basis for human autonomic rhythms. Annals of Medicine, 32: 341349.Google Scholar
Eckberg, D. L. (2003). The human respiratory gate. Journal of Physiology, 548: 339352.Google Scholar
Fabiani, M., Low, K. A., Tan, C. H., Zimmerman, B., Fletcher, M. A., Schneider-Garces, N., … & Gratton, G. 2014. Taking the pulse of aging: mapping pulse pressure and elasticity in cerebral arteries with optical methods. Psychophysiology, 51: 10721088.Google Scholar
Fitzsimons, J. T. (1998). Angiotensin, thirst, and sodium appetite. Physiological Reviews, 78: 583686.Google Scholar
Frederiks, J., Swenne, C. A., TenVoorde, B. J., Honzíková, N., Levert, J. V., Maan, A. C., … & Bruschke, A. V. (2000). The importance of high-frequency paced breathing in spectral baroreflex sensitivity assessment. Journal of Hypertension, 18: 16351644.Google Scholar
Fukuda, N. & Granzier, H. L. (2005). Titin/connectin-based modulation of the Frank-Starling mechanism of the heart. Journal of Muscle Research & Cell Motility, 26: 319323.Google Scholar
Fukuda, N., Terui, T., Ishiwata, S. I., & Kurihara, S. (2010). Titin-based regulations of diastolic and systolic functions of mammalian cardiac muscle. Journal of Molecular and Cellular Cardiology, 48: 876881.Google Scholar
Fukuda, N., Terui, T., Ohtsuki, I., Ishiwata, S. I., & Kurihara, S. (2009). Titin and troponin: central players in the Frank-Starling mechanism of the heart. Current Cardiology Reviews, 5: 119124.Google Scholar
Furedy, J. J., Heslegrave, R. J., & Scher, H. (1992). T-wave amplitude utility revisited: some physiological and psychophysiological considerations. Biological Psychology, 33: 241248.Google Scholar
Gang, Y. & Malik, M. (2002). Heart rate variability in critical care medicine. Current Opinion in Critical Care, 8: 371375.Google Scholar
Gianaros, P. J., May, J. C., Siegle, G. J., & Jennings, J. R. (2005). Is there a functional neural correlate of individual differences in cardiovascular reactivity? Psychosomatic Medicine, 67: 3139.Google Scholar
Gianaros, P. J., Onyewuenyi, I. C., Sheu, L. K., Christie, I. C., & Critchley, H. D. (2012). Brain systems for baroreflex suppression during stress in humans. Human Brain Mapping, 33: 17001716.Google Scholar
Gianaros, P. J. & Quigley, K. S. (2001). Autonomic origins of a nonsignal stimulus-elicited bradycardia and its habituation in humans. Psychophysiology, 38: 540547.Google Scholar
Gianaros, P. J., Van Der Veen, F. M., & Jennings, J. R. (2004). Regional cerebral blood flow correlates with heart period and high frequency heart period variability during working memory tasks: implications for the cortical and subcortical regulation of cardiac autonomic activity. Psychophysiology, 41: 521530.Google Scholar
Gibbons, R. J. & Araoz, P. A. (2004). The year in cardiac imaging. Journal of the American College of Cardiology, 44: 19371944.Google Scholar
Glaser, R., Kiecolt-Glaser, J. K., Malarkey, W. B., & Sheridan, J. F. (1998). The influence of psychological stress on the immune response to vaccines. Annals of the New York Academy of Sciences, 840: 649655.Google Scholar
Goedhart, A. D., Kupper, N., Willemsen, G., Boomsma, D. I., & de Geus, E. J. (2006). Temporal stability of ambulatory stroke volume and cardiac output measured by impedance cardiography. Biological Psychology, 72: 110117.Google Scholar
Goedhart, A. D., Willemsen, G., Houtveen, J. H., Boomsma, D. I., & De Geus, E. J. (2008). Comparing low frequency heart rate variability and preejection period: two sides of a different coin. Psychophysiology, 45: 10861090.Google Scholar
Goldberger, A. L. (2013). Clinical Electrocardiography: A Simplified Approach, 8th edn. Philadelphia, PA: Elsevier Saunders.Google Scholar
Goldin, J. G., Ratib, O., & Aberle, D. R. (2000). Contemporary cardiac imaging: an overview. Journal of Thoracic Imaging, 15: 218229.Google Scholar
Goldstein, D. S., Bentho, O., Park, M. Y., & Sharabi, Y. (2011). Low-frequency power of heart rate variability is not a measure of cardiac sympathetic tone but may be a measure of modulation of cardiac autonomic outflows by baroreflexes. Experimental Physiology, 96: 12551261.Google Scholar
Graham, F. K. (1978). Constraints on measuring heart rate and period sequentially through real and cardiac time. Psychophysiology, 15: 492495.Google Scholar
Gratton, G. & Fabiani, M. (2010). Fast optical imaging of human brain function. Frontiers in Human Neuroscience, 4: 52.Google Scholar
Gray, A. L., Johnson, T. A., Ardell, J. L., & Massari, V. J. (2004a). Parasympathetic control of the heart: II. A novel interganglionic intrinsic cardiac circuit mediates neural control of heart rate. Journal of Applied Physiology, 96: 22732278.Google Scholar
Gray, A. L., Johnson, T. A., Lauenstein, J. M., Newton, S. S., Ardell, J. L., & Massari, V. J. (2004b). Parasympathetic control of the heart: III. Neuropeptide Y-immunoreactive nerve terminals synapse on three populations of negative chronotropic vagal preganglionic neurons. Journal of Applied Physiology, 96: 22792287.Google Scholar
Gray, M., Nagai, Y., & Critchley, H. D. (2012). Brain imaging of stress and cardiovascular responses. In Hjemdahl, P., Rosengren, A., & Steptoe, A. (eds.), Stress and Cardiovascular Disease (pp. 129148). London: Springer.Google Scholar
Gray, M. A., Rylander, K., Harrison, N. A., Wallin, B. G., & Critchley, H. D. (2009). Following one’s heart: cardiac rhythms gate central initiation of sympathetic reflexes. Journal of Neuroscience, 29: 18171825.Google Scholar
Grisk, O. & Rettig, R. (2004). Interactions between the sympathetic nervous system and the kidneys in arterial hypertension. Cardiovascular Research, 61: 238246.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: 201216.Google Scholar
Grossman, P. & Kollai, M. (1993). Respiratory sinus arrhythmia, cardiac vagal tone, and respiration: within and between individual relations. Psychophysiology, 30: 486495.Google Scholar
Grossman, P., van Beek, J., & Wientjes, C. (1990). A comparison of three quantification methods for estimation of respiratory sinus arrhythmia. Psychophysiology, 27: 702714.Google Scholar
Grossman, P., Wilhelm, F. H., & Spoerle, M. (2004). Respiratory sinus arrhythmia, cardiac vagal control, and daily activity. American Journal of Physiology: Heart & Circulatory Physiology, 287: H728H734.Google Scholar
Guimaraes, S. & Moura, D. (2001). Vascular adrenoceptors: an update. Pharmacological Review, 53: 319356.Google Scholar
Guthrie, D. & Yucha, C. (2004). Urinary concentration and dilution. Nephrolology Nursing Journal, 31: 297303.Google Scholar
Guyton, A. C. & Hall, J. E. (2010). Textbook of Medical Physiology, 12th edn. Philadelphia: W. B. Saunders.Google Scholar
Hall, J. E. (2010). Guyton and Hall Textbook of Medical Physiology. New York: Elsevier.Google Scholar
Hansson, G. K. & Hermansson, A. (2011). The immune system in atherosclerosis. Nature Immunology, 12: 204212.Google Scholar
Hawkley, L. C., Burleson, M. H., Berntson, G. G., & Cacioppo, J. T. (2003). Loneliness in everyday life: cardiovascular activity, psychosocial context, and health behaviors. Journal of Personality and Social Psychology, 85: 105120.Google Scholar
Henelius, A., Sallinen, M., Huotilainen, M., Müller, K., Virkkala, J., & Puolamäki, K. (2014). Heart rate variability for evaluating vigilant attention in partial chronic sleep restriction. Sleep, 37: 12571267.Google Scholar
Henry, I. C., Bernstein, D. P., & Banet, M. J. (2012). Stroke volume obtained from the brachial artery using transbrachial electrical bioimpedance velocimetry. In Conference Proceedings of the IEEE Engineering Medicine Biology Society, 2012 (pp. 142145). Piscataway, NJ: IEEE.Google Scholar
Higgins, C. B. (2000). Cardiac imaging. Radiology, 217: 410.Google Scholar
Hoetink, A. E., Faes, T. J., Schuur, E. H., Gorkink, R., Goovaerts, H. G., Meijer, J. H., & Heethaar, R. M. (2002). Comparing spot electrode arrangements for electric impedance cardiography. Physiological Measurement, 23: 457467.Google Scholar
Hoetink, A. E., Faes, T. J., Visser, K. R., & Heethaar, R. M. (2004). On the flow dependency of the electrical conductivity of blood. IEEE Transactions on Biomedical Engineering, 51: 12511261.Google Scholar
Ikarashi, A., Nogawa, M., Yamakoshi, T., Tanaka, S., & Yamakoshi, K. (2006). An optimal spot-electrodes array for electrical impedance cardiography through determination of impedance mapping of a regional area along the medial line on the thorax. Conference Proceedings IEEE Engineering in Medicine and Biology Society, 1: 32023205.Google Scholar
Iwata, J. & LeDoux, J. E. (1988). Dissociation of associative and nonassociative concomitants of classical fear conditioning in the freely behaving rat. Behavioral Neuroscience, 102: 6676.Google Scholar
Jagadeesh, A. M., Singh, N. G., & Mahankali, S. (2012). A comparison of a continuous noninvasive arterial pressure (CNAP™) monitor with an invasive arterial blood pressure monitor in the cardiac surgical ICU. Annals of Cardiac Anaesthesia, 15: 180184.Google Scholar
Jennings, J. R., Kamarck, T. W., Everson Rose, S. A., Kaplan, G. A., Manuck, S. B., & Salonen, J. T. (2004). Exaggerated blood pressure responses during mental stress are prospectively related to enhanced carotid atherosclerosis in middle-aged Finnish men. Circulation, 110: 21982203.Google Scholar
Jennings, J. R., Tahmoush, A. J., & Redmond, D. P. (1980). Noninvasive measurement of peripheral vascular activity. In Martin, I. & Venables, P. H. (eds.), Techniques in Psychophysiology (pp. 69137). New York: John Wiley.Google Scholar
Johnson, T. A., Gray, A. L., Lauenstein, J. M., Newton, S. S., & Massari, V. J. (2004). Parasympathetic control of the heart: I. An interventriculoseptal ganglion is the major source of the vagal intracardiac innervation of the ventricles. Journal of Applied Physiology, 96: 22652272.Google Scholar
Joyner, M. J. & Casey, D. P. (2015). Regulation of increased blood flow (hyperemia) to muscles during exercise: a hierarchy of competing physiological needs. Physiological Reviews, 95: 549601.Google Scholar
Karelina, K., Norman, G. J., Zhang, N., Morris, J. S., Peng, H., & DeVries, A. C. (2009). Social isolation alters neuroinflammatory response to stroke. Proceedings of the National Academy of Sciences of the USA, 106: 58955900.Google Scholar
Kauppinen, P. K., Hyttinen, J. A., & Malmivuo, J. A. (1998). Sensitivity distributions of impedance cardiography using band and spot electrodes analyzed by a three-dimensional computer model. Annals of Biomedical Engineering, 26: 694702.Google Scholar
Kauppinen, P. K., Koobi, T., Hyttinen, J., & Malmivuo, J. (2000). Segmental composition of whole body impedance cardiogram estimated by computer simulations and clinical experiments. Clinical Physiology, 20: 106113.Google Scholar
Kelsey, R. M., Reiff, S., Wiens, S., Schneider, T. R., Mezzacappa, E. S., & Guethlein, W. (1998). The ensemble-averaged impedance cardiogram: an evaluation of scoring methods and interrater reliability. Psychophysiology, 35: 337340.Google Scholar
Kemmotsu, O., Ueda, M., Otsuka, H., Yamamura, T., Winter, D. C., & Eckerle, J. S. (1991). Arterial tonometry for noninvasive, continuous blood pressure monitoring during anesthesia. Anesthesiology, 75: 333340.Google Scholar
Kline, K. P., Ginsburg, G. P., & Johnston, J. R. (1998). T-wave amplitude: relationships to phasic RSA and heart period changes. International Journal of Psychophysiology, 29: 291301.Google Scholar
Koh, J., Brown, T. E., Beightol, L. A., & Eckberg, D. L. (1998). Contributions of tidal lung inflation to human R-R interval and arterial pressure fluctuations. Journal of the Autonomic Nervous System, 68: 8995.Google Scholar
Kreibig, S. D., Gendolla, G. H., & Scherer, K. R. (2012). Goal relevance and goal conduciveness appraisals lead to differential autonomic reactivity in emotional responding to performance feedback. Biological Psychology, 91: 365375.Google Scholar
Kubicek, W. G., Karnegis, J. N., Patterson, R. P., Witsoe, D. A., & Mattson, R. H. (1966). Development and evaluation of an impedance cardiac output system. Aerospace Medicine, 37: 12081212.Google Scholar
Kurzen, H. & Schallreuter, K. U. (2004). Novel aspects in cutaneous biology of acetylcholine synthesis and acetylcholine receptors. Experimental Dermatology, 13: 2730.Google Scholar
Kuvin, J. T., Patel, A. R., Sliney, K. A., Pandian, N. G., Rand, W. M., Udelson, J. E., & Karas, R. H. (2001). Peripheral vascular endothelial function testing as a noninvasive indicator of coronary artery disease. Journal of the American College of Cardiology, 38: 18431849.Google Scholar
Kuvin, J. T., Patel, A. R., Sliney, K. A., Pandian, N. G., Sheffy, J., Schnall, R. P., … & Udelson, J. E. (2003). Assessment of peripheral vascular endothelial function with finger arterial pulse wave amplitude. American Heart Journal, 146: 168174.Google Scholar
Lacey, J. I. & Lacey, B. C. (1962). The law of initial value in the longitudinal study of autonomic constitution: reproducibility of autonomic responses and response patterns over a four year interval. Annals of the New York Academy of Sciences, 98: 12571290.Google Scholar
Landis, S. C. (1996). The development of cholinergic sympathetic neurons: a role for neuropoietic cytokines? Perspectives in Developmental Neurobiology, 4: 5363.Google Scholar
Lane, R. D., Reiman, E. M., Ahern, G. L., & Thayer, J. F. (2001). Activity in medial prefrontal cortex correlates with vagal component of heart rate variability during emotion. Brain and Cognition, 47: 97100.Google Scholar
Lang, P. J. (2014). Emotion’s response patterns: the brain and the autonomic nervous system. Emotion Review, 6: 9399.Google Scholar
Laude, D., Elghozi, J. L., Girard, A., Bellard, E., Bouhaddi, M., Castiglioni, P., … & Stauss, H. M. (2004). Comparison of various techniques used to estimate spontaneous baroreflex sensitivity (the EuroBaVar study). American Journal of Physiology: Regulatory, Integrative and Comparative Physiolology, 286: R226R231.Google Scholar
Lehrer, P. M., Vaschillo, E., Vaschillo, B., Lu, S. E., Eckberg, D. L., Edelberg, R., … & Hamer, R. M. (2003). Heart rate variability biofeedback increases baroreflex gain and peak expiratory flow. Psychosomatic Medicine, 65: 796805.Google Scholar
Levenson, R. W. (2014). Emotion and the autonomic nervous system: introduction to the special section. Emotion Review, 6: 9192.Google Scholar
Levy, M. N. (1984). Cardiac sympathetic–parasympathetic interactions. Federation Proceedings, 43: 25982602.Google Scholar
Libby, P. (2003). Vascular biology of atherosclerosis: overview and state of the art. American Journal of Cardiology, 91: 3A6A.Google Scholar
Lindh, B. & Hokfelt, T. (1990). Structural and functional aspects of acetylcholine peptide coexistence in the autonomic nervous system. Progress in Brain Research, 84: 175191.Google Scholar
Litvack, D. A., Oberlander, T. F., Carney, L. H., & Saul, J. P. (1995). Time and frequency domain methods for heart rate variability analysis: a methodological comparison. Psychophysiology, 32: 492504.Google Scholar
Llabre, M. M., Ironson, G. H., Spitzer, S. B., Gellman, M. D., Weidler, D. J., & Schneiderman, N. (1988). How many blood pressure measurements are enough? An application of generalizability theory to the study of blood pressure reliability. Psychophysiology, 25: 97106.Google Scholar
Longmore, J., Bradshaw, C. M., & Szabadi, E. (1985). Effects of locally and systemically administered cholinoceptor antagonists on the secretory response of human eccrine sweat glands to carbachol. British Journal of Clinical Pharmacology, 20: 17.Google Scholar
Lozano, D. L., Norman, G., Knox, D., Wood, B. L., Miller, B. D., Emery, C. F., & Berntson, G. G. (2007). Where to B in dZ/dt. Psychophysiology, 44: 113119.Google Scholar
Luchner, A. & Schunkert, H. (2004). Interactions between the sympathetic nervous system and the cardiac natriuretic peptide system. Cardiovascular Research, 63: 443449.Google Scholar
Lymperopoulos, A. (2013). Physiology and pharmacology of the cardiovascular adrenergic system. Frontiers in Physiology, 4: 240.Google Scholar
Macfarlane, P. W., van Oosterom, A., Janse, M., Kligfield, P., Camm, J., & Pahlm, O. (2012). Basic Electrocardiology: Cardiac Electrophysiology, ECG Systems and Mathematical Modeling. New York: Springer.Google Scholar
Machado-Moreira, C. A., McLennan, P. L., Lillioja, S., van Dijk, W., Caldwell, J. N., & Taylor, N. A. (2012). The cholinergic blockade of both thermally and non-thermally induced human eccrine sweating. Experimental Physiology, 97: 930942.Google Scholar
Malliani, A. (1999). The pattern of sympathovagal balance explored in the frequency domain. News in Physiological Sciences, 14: 111117.Google Scholar
Martínez-García, P., Lerma, C., & Infante, O. (2012). Baroreflex sensitivity estimation by the sequence method with delayed signals. Clinical Autonomic Research, 22: 289297.Google Scholar
Matthews, K. A., Salomon, K., Brady, S. S., & Allen, M. T. (2003). Cardiovascular reactivity to stress predicts future blood pressure in adolescence. Psychosomatic Medicine, 65: 410415.Google Scholar
Matthews, S. C., Paulus, M. P., Simmons, A. N., Nelesen, R. A., & Dimsdale, J. E. (2004). Functional subdivisions within anterior cingulate cortex and their relationship to autonomic nervous system function. NeuroImage, 22: 11511156.Google Scholar
Monti, A., Medigue, C., & Mangin, L. (2002). Instantaneous parameter estimation in cardiovascular time series by harmonic and time-frequency analysis. IEEE Transactions in Biomedical Engineering, 49: 15471556.Google Scholar
Moshkovitz, Y., Kaluski, E., Milo, O., Vered, Z., & Cotter, G. (2004). Recent developments in cardiac output determination by bioimpedance: comparison with invasive cardiac output and potential cardiovascular applications. Current Opinion in Cardiology, 19: 229237.Google Scholar
Norman, G. J., Berntson, G. G., & Cacioppo, J. T. (2014). Emotion, somatovisceral afference, and autonomic regulation. Emotion Review, 6: 113123.Google Scholar
Norman, G. J., Karelina, K., Morris, J. S., Zhang, N., Cochran, M., & DeVries, A. C. (2010a). Social interaction prevents the development of depressive-like behavior post nerve injury in mice: a potential role for oxytocin. Psychosomatic Medicine, 72: 519526.Google Scholar
Norman, G. J., Zhang, N., Morris, J. S., Karelina, K., Berntson, G. G., & DeVries, A. C. (2010b). Social interaction modulates autonomic, inflammatory, and depressive-like responses to cardiac arrest and cardiopulmonary resuscitation. Proceedings of the National Academy of Sciences of the USA, 107: 1634216347.Google Scholar
O’Brien, E. (1996). Review. A century of confusion: which bladder for accurate blood pressure measurement? Journal of Human Hypertension, 10: 565572.Google Scholar
Padgett, D. A., Sheridan, J. F., Dorne, J., Berntson, G. G., Candelora, J., & Glaser, R. (1998). Social stress and the reactivation of latent herpes simplex virus type 1. Proceedings of the National Academy of Sciences of the USA, 95: 72317235.Google Scholar
Parati, G., Di Rienzo, M., & Mancia, G. (2000). How to measure baroreflex sensitivity: from the cardiovascular laboratory to daily life. Journal of Hypertension, 18: 719.Google Scholar
Parati, G., Ongaro, G., Bilo, G., Glavina, F., Castiglioni, P., Di Rienzo, M., & Mancia, G. (2003). Noninvasive beat to beat blood pressure monitoring: new developments. Blood Pressure Monitoring, 8: 3136.Google Scholar
Park, G. & Thayer, J. F. (2014). From the heart to the mind: cardiac vagal tone modulates top-down and bottom-up visual perception and attention to emotional stimuli. Frontiers in Psychology, 5: 278.Google Scholar
Parker, P., Celler, B. G., Potter, E. K., & McCloskey, D. I. (1984). Vagal stimulation and cardiac slowing. Journal of the Autonomic Nervous System, 11: 226231.Google Scholar
Parry, M. J. & McFetridge-Durdle, J. (2006). Ambulatory impedance cardiography: a systematic review. Nursing Research, 55: 283291.Google Scholar
Pépin, J. L., Tamisier, R., Borel, J. C., Baguet, J. P., & Lévy, P. (2009). A critical review of peripheral arterial tone and pulse transit time as indirect diagnostic methods for detecting sleep disordered breathing and characterizing sleep structure. Current Opinion in Pulmonary Medicine, 15: 550558.Google Scholar
Persson, P. B., Di Rienzo, M., Castiglioni, P., Cerutti, C., Pagani, M., Honzikova, N., … & Parati, G. (2001). Time versus frequency domain techniques for assessing baroreflex sensitivity. Journal of Hypertension, 19: 16991705.Google Scholar
Picciotto, M. R., Higley, M. J., & Mineur, Y. S. (2012). Acetylcholine as a neuromodulator: cholinergic signaling shapes nervous system function and behavior. Neuron, 76: 116129.Google Scholar
Pirola, F. T. & Potter, E. K. (1990). Vagal action on atrioventricular conduction and its inhibition by sympathetic stimulation and neuropeptide Y in anaesthetised dogs. Journal of the Autonomic Nervous System, 31: 112.Google Scholar
Pittman, S. D., Ayas, N. T., MacDonald, M. M., Malhotra, A., Fogel, R. B., & White, D. P. (2004). Using a wrist-worn device based on peripheral arterial tonometry to diagnose obstructive sleep apnea: in-laboratory and ambulatory validation. Sleep, 27: 923933.Google Scholar
Poliakova, N., Dionne, G., Dubreuil, E., Ditto, B., Pihl, R. O., Pérusse, D., … & Boivin, M. (2014). A methodological comparison of the Porges algorithm, fast Fourier transform, and autoregressive spectral analysis for the estimation of heart rate variability in 5-month-old infants. Psychophysiology, 51: 579583.Google Scholar
Porges, S. W. (1992). Autonomic regulation and attention. In Campbell, B. A., Hayne, H., & Richardson, R. (eds.), Attention and Information Processing in Infants and Adults (pp. 201223). Hillsdale, NJ: Lawrence Erlbaum Associates.Google Scholar
Porges, S. W. & Bohrer, R. E. (1990). Analysis of periodic processes in psychophysiological research. In Cacioppo, J. T. & Tassinary, L. G. (eds.), Principles of Psychophysiology: Physical, Social and Inferential Elements (pp. 708753). Cambridge University Press.Google Scholar
Powell, N. D., Sloan, E. K., Bailey, M. T., Arevalo, J. M., Miller, G. E., Chen, E., & Cole, S. W. (2013). Social stress up-regulates inflammatory gene expression in the leukocyte transcriptome via β-adrenergic induction of myelopoiesis. Proceedings of the National Academy of Sciences of the USA, 110: 1657416579.Google Scholar
Pumprla, J., Howorka, K., Groves, D., Chester, M., & Nolan, J. (2002). Functional assessment of heart rate variability: physiological basis and practical applications. International Journal of Cardiology, 84: 114.Google Scholar
Qu, M. H., Zhang, Y. J., Webster, J. G., & Tompkins, W. J. (1986). Motion artifact from spot and band electrodes during impedance cardiography. IEEE Transactions in Biomedical Engineering, 33: 10291036.Google Scholar
Quan, N., Avitsur, R., Stark, J. L., He, L., Lai, W., Dhabhar, F., & Sheridan, J. F. (2003). Molecular mechanisms of glucocorticoid resistance in splenocytes of socially stressed male mice. Journal of Neuroimmunology, 137: 5158.Google Scholar
Quigley, K. S. & Berntson, G. G. (1996) Autonomic interactions and chronotropic control of the heart: heart period vs. heart rate. Psychophysiology, 33: 605611.Google Scholar
Quigley, K. S. & Stifter, C. A. (2006). A comparative validation of sympathetic reactivity in children and adults. Psychophysiology, 43: 357365.Google Scholar
Raaijmakers, E., Faes, T. J., Scholten, R. J., Goovaerts, H. G., & Heethaar, R. M. (1999). A meta-analysis of published studies concerning the validity of thoracic impedance cardiography. Annals of the New York Academy of Sciences, 873: 121127.Google Scholar
Randall, W., Wurster, R., Randall, D., & Xi Moy, S. (1996). From cardioaccelerator and inhibitory nerves to a “heart brain”: an evolution of concepts. In Shepard, J. T. & Vatner, S. F. (eds.), Nervous Control of the Heart. Amsterdam: Harwood Academic Publishers.Google Scholar
Rashba, E. J., Cooklin, M., MacMurdy, K., Kavesh, N., Kirk, M., Sarang, S., … & Gold, M. R. (2002). Effects of selective autonomic blockade on T-wave alternans in humans. Circulation, 105: 837842.Google Scholar
Ren, L. M., Furukawa, Y., Karasawa, Y., Murakami, M., Takei, M., Narita, M., & Chiba, S. (1991). Differential inhibition of neuropeptide Y on the chronotropic and inotropic responses to sympathetic and parasympathetic stimulation in the isolated, perfused dog atrium. Journal of Pharmacology and Experimental Therapeutics, 259: 3843.Google Scholar
Reyes del Paso, G. A., González, I., & Hernández, J. A. (2004a). Baroreceptor sensitivity and effectiveness varies differentially as a function of cognitive-attentional demands. Biological Psychology, 67: 385395.Google Scholar
Reyes del Paso, G. A., Hernández, J. A., & González, I. (2004b). Differential analysis in the time domain of the baroreceptor cardiac reflex sensitivity as a function of sequence length. Psychophysiology, 41: 483488.Google Scholar
Reyes del Paso, G. A., Langewitz, W., Mulder, L. J., van Roon, A., & Duschek, S. (2013). The utility of low frequency heart rate variability as an index of sympathetic cardiac tone: a review with emphasis on a reanalysis of previous studies. Psychophysiology, 50: 477487.Google Scholar
Richardson, R. J., Grkovic, I., & Anderson, C. R. (2003). Immunohistochemical analysis of intracardiac ganglia of the rat heart. Cell and Tissue Research, 314: 337350.Google Scholar
Riese, H., Groot, P. F., van den Berg, M., Kupper, N. H., Magnee, E. H., Rohaan, E. J., … & de Geus, E. J. (2003). Large-scale ensemble averaging of ambulatory impedance cardiograms. Behavioral Research Methods, Instruments and Computers, 35: 467477.Google Scholar
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: 613621.Google Scholar
Rose, S. C. (2000). Noninvasive vascular laboratory for evaluation of peripheral arterial occlusive disease: Part I. Hemodynamic principles and tools of the trade. Journal of Vascular and Interventional Radiology, 11: 11071114.Google Scholar
Rosengren, A., Hawken, S., Ounpuu, S., Sliwa, K., Zubaid, M., Almahmeed, W. A., … & Yusuf, S. (2004). Association of psychosocial risk factors with risk of acute myocardial infarction in 11119 cases and 13648 controls from 52 countries (the INTERHEART study): case control study. Lancet, 364: 953962.Google Scholar
Sampaio, K. N., Mauad, H., Spyer, K. M., & Ford, T. W. (2003). Differential chronotropic and dromotropic responses to focal stimulation of cardiac vagal ganglia in the rat. Experimental Physiology, 88: 315327.Google Scholar
Shapiro, D., Jamner, L. D., Lane, J. D., Light, K. C., Myrtek, M., Sawada, Y., & Steptoe, A. (1996). Blood pressure publication guidelines. Psychophysiology, 33: 112.Google Scholar
Shechter, M., Issachar, A., Marai, I., Koren-Morag, N., Freinark, D., Shahar, Y., & Feinberg, M. S. (2009). Long-term association of brachial artery flow-mediated vasodilation and cardiovascular events in middle-aged subjects with no apparent heart disease. International Journal of Cardiology, 134: 5258.Google Scholar
Sheridan, J. F., Stark, J. L., Avitsur, R., & Padgett, D. A. (2000). Social disruption, immunity, and susceptibility to viral infection: role of glucocorticoid insensitivity and NGF. Annals of the New York Academy of Sciences, 917: 894905.Google Scholar
Sherwood, A., Allen, M. T., Fahrenberg, J., Kelsey, R. M., Lovallo, W. R., & van Doornen, L. J. (1990). Methodological guidelines for impedance cardiography. Psychophysiology, 27: 123.Google Scholar
Sherwood, A., Dolan, C. A., & Light, K. C. (1990). Hemodynamics of blood pressure responses during active and passive coping. Psychophysiology, 27: 656668.Google Scholar
Sherwood, A., Royal, S. A., Hutcheson, J. S., & Turner, J. R. (1992). Comparison of impedance cardiographic measurements using band and spot electrodes. Psychophysiology, 29: 734741.Google Scholar
Simmons, W. K., Avery, J. A., Barcalow, J. C., Bodurka, J., Drevets, W. C., & Bellgowan, P. (2013). Keeping the body in mind: insula functional organization and functional connectivity integrate interoceptive, exteroceptive, and emotional awareness. Human Brain Mapping, 34: 29442958.Google Scholar
Smith, L. L., Kukielka, M., & Billman, G. E. (2005). Heart rate recovery after exercise: a predictor of ventricular fibrillation susceptibility after myocardial infarction. American Journal of Physiology: Heart and Circulatory Physiology, 288: H17631769.Google Scholar
Somsen, R. J., Jennings, J. R., & Van der Molen, M. W. (2004). The cardiac cycle time effect revisited: temporal dynamics of the central-vagal modulation of heart rate in human reaction time tasks. Psychophysiology, 41: 941953.Google Scholar
Stankovic, Z., Allen, B. D., Garcia, J., Jarvis, K. B., & Mark, M. (2014). 4D flow imaging with MRI. Cardiovascular Diagnosis and Therapy, 4: 173192.Google Scholar
Steptoe, A., Godaert, G., Ross, A., & Schreurs, P. (1983). The cardiac and vascular components of pulse transmission time: a computer analysis of systolic time intervals. Psychophysiology, 20: 251259.Google Scholar
Steptoe, A. & Sawada, Y. (1989). Assessment of baroreceptor reflex function during mental stress and relaxation. Psychophysiology, 26: 140147.Google Scholar
Strike, P. C. & Steptoe, A. (2004). Psychosocial factors in the development of coronary artery disease. Progress in Cardiovascular Disesase, 46: 337347.Google Scholar
Stuiver, A., de Waard, D., Brookhuis, K. A., Dijksterhuis, C., Lewis-Evans, B., & Mulder, L. J. (2012). Short-term cardiovascular responses to changing task demands. International Journal of Psychophysiology, 85: 153160.Google Scholar
Stuiver, A. & Mulder, B. (2014). Cardiovascular state changes in simulated work environments. Frontiers in Neuroscience, 8: article 399.Google Scholar
Swenne, C. A. (2013). Baroreflex sensitivity: mechanisms and measurement. Netherlands Heart Journal, 21: 5860.Google Scholar
Takahashi, H., Maehara, K., Onuki, N., Saito, T., & Maruyama, Y. (2003). Decreased contractility of the left ventricle is induced by the neurotransmitter acetylcholine, but not by vagal stimulation in rats. Japanese Heart Journal, 44: 257270.Google Scholar
Task Force of the European Society of Cardiology and the North American Society of Pacing and Electrophysiology (1996). Heart rate variability: standards of measurement, physiological interpretation, and clinical use. Circulation, 93: 10431065.Google Scholar
Ter Horst, G. J., Hautvast, R. W., De Jongste, M. J., & Korf, J. (1996). Neuroanatomy of cardiac activity regulating circuitry: a transneuronal retrograde viral labelling study in the rat. European Journal of Neuroscience, 8: 20292041.Google Scholar
Thayer, J. F. & Lane, R. D. (2009). Claude Bernard and the heart–brain connection: Further elaboration of a model of neurovisceral integration. Neuroscience & Biobehavioral Reviews, 33: 8188.Google Scholar
Thayer, J. F. & Uijtdehaage, S. H. (2001). Derivation of chronotropic indices of autonomic nervous system activity using impedance cardiography. Biomedical Sciences Instrumentation, 37: 331336.Google Scholar
Thayer, J. F., Yamamoto, S. S., & Brosschot, J. F. (2010). The relationship of autonomic imbalance, heart rate variability and cardiovascular disease risk factors. International Journal of Cardiology, 141: 122131.Google Scholar
Thijssen, D. H., Black, M. A., Pyke, K. E., Padilla, J., Atkinson, G., Harris, R. A., … & Green, D. J. (2011). Assessment of flow-mediated dilation in humans: a methodological and physiological guideline. American Journal of Physiology: Heart and Circulatory Physiology, 300: H2H12.Google Scholar
Tomaka, J., Blascovich, J., Kelsey, R. M., & Leitten, C. L. (1993). Subjective, physiological, and behavioral effects of threat and challenge appraisal. Journal of Personality and Social Psychology, 65: 248260.Google Scholar
Tomaka, J., Blascovich, J., & Swart, L. (1994). Effects of vocalization on cardiovascular and electrodermal responses during mental arithmetic. International Journal of Psychophysiology, 18: 2333.Google Scholar
Ursino, M. & Magosso, E. (2003). Short-term autonomic control of cardiovascular function: a mini review with the help of mathematical models. Journal of Integrative Neuroscience, 2: 219247.Google Scholar
Vallbo, A. B., Hagbarth, K. E., & Wallin, B. G. (2004). Microneurography: how the technique developed and its role in the investigation of the sympathetic nervous system. Journal of Applied Physiology, 96: 12621269.Google Scholar
Van De Water, J. M., Miller, T. W., Vogel, R. L., Mount, B. E., & Dalton, M. L. (2003). Impedance cardiography: the next vital sign technology? Chest, 123: 20282033.Google Scholar
van der Meer, B. J., Vonk Noordegraaf, A., Bax, J. J., Kamp, O., & de Vries, P. M. (1999). Non-invasive evaluation of left ventricular function by means of impedance cardiography. Acta Anaesthesiology Scandinavica, 43: 130134.Google Scholar
van Dijk, A. E., van Lien, R., van Eijsden, M., Gemke, R. J., Vrijkotte, T. G., & de Geus, E. J. (2013). Measuring cardiac autonomic nervous system (ANS) activity in children. Journal of Visualized Experiments (JOVE), 29: e50073. www.ncbi.nlm.nih.gov/pmc/articles/PMC3667644/Google Scholar
van Lien, R., Neijts, M., Willemsen, G., & de Geus, E. J. (2015). Ambulatory measurement of the ECG T-wave amplitude. Psychophysiology, 52: 225237.Google Scholar
van Lien, R., Schutte, N. M., Meijer, J. H., & de Geus, E. J. (2013). Estimated preejection period (PEP) based on the detection of the R-wave and dZ/dt-min peaks does not adequately reflect the actual PEP across a wide range of laboratory and ambulatory conditions. International Journal of Psychophysiology, 87: 6069.Google Scholar
van Montfrans, G. A. (2001). Oscillometric blood pressure measurements: progress and problems. Blood Pressure Monitoring, 6: 287290.Google Scholar
Van Roon, A. M., Mulder, L. J., Althaus, M., & Mulder, G. (2004). Introducing a baroreflex model for studying cardiovascular effects of mental workload. Psychophysiology, 41: 961981.Google Scholar
van Vark, L. C., Bertrand, M., Akkerhuis, K. M., Brugts, J. J., Fox, K., Mourad, J. J., & Boersma, E. (2012). Angiotensin-converting enzyme inhibitors reduce mortality in hypertension: a meta-analysis of randomized clinical trials of renin–angiotensin–aldosterone system inhibitors involving 158998 patients. European Heart Journal, 33: 20882097.Google Scholar
Vuurmans, T. J. L., Boer, P., & Koomans, H. A. (2003). Effects of endothelin1 and endothelin1 receptor blockade on cardiac output, aortic pressure, and pulse wave velocity in humans. Hypertension, 41: 12531258.Google Scholar
Wallin, B. G. & Charkoudian, N. (2007). Sympathetic neural control of integrated cardiovascular function: insights from measurement of human sympathetic nerve activity. Muscle and Nerve, 36: 595614.Google Scholar
Wang, Y. P., Kuo, T. B., Lai, C. T., Lee, G. S., & Yang, C. C. (2012). Effects of breathing frequency on baroreflex effectiveness index and spontaneous baroreflex sensitivity derived by sequence analysis. Journal of Hypertension, 30: 21512158.Google Scholar
Ward, A. R., Alarcón, G., Nigg, J. T., & Musser, E. D. (2015). Variation in parasympathetic dysregulation moderates short-term memory problems in childhood attention-deficit/hyperactivity disorder. Journal of Abnormal Child Psychology, 43: 15731583.Google Scholar
Watkins, L., Fainman, C., Dimsdale, J., & Ziegler, M. (1995). Assessment of baroreflex control from beat-to-beat blood pressure and heart rate changes: a validation study. Psychophysiology, 32: 411414.Google Scholar
Weber, E. J., Molenaar, P. C., & van der Molen, M. W. (1992). A nonstationarity test for the spectral analysis of physiological time series with an application to respiratory sinus arrhythmia. Psychophysiology, 29: 5565.Google Scholar
Weyman, A. E. (2005). The year in echocardiography. Journal of the American College of Cardiology, 45: 448455.Google 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 Sciences Instrumentation, 40: 317324.Google Scholar
Wilhelm, F. H., Grossman, P., & Roth, W. T. (1999). Analysis of cardiovascular regulation. Biomedical Sciences and Instrumentation, 35: 135140.Google Scholar
Wilkinson, I. B. & Webb, D. J. (2001). Venous occlusion plethysmography in cardiovascular research: methodology and clinical applications. British Journal of Clinical Pharmacology, 52: 631646.Google Scholar
Wood, D. (2001). Established and emerging cardiovascular risk factors. American Heart Journal, 141: 4957.Google Scholar
Woods, R. L. (2004). Cardioprotective functions of atrial natriuretic peptide and B-type natriuretic peptide: a brief review. Clinical and Experimental Pharmacology and Physiology, 31: 791794.Google Scholar

References

Andreassi, J. L. (2007). Psychophysiology: Human Behavior and Physiological Response, 5th edn. Hillsdale, NJ: Lawrence Erlbaum Associates.Google Scholar
Aue, T., Hoeppli, M., Piguet, C., Sterpenich, V., & Vuilleumier, P. (2013). Visual avoidance in phobia: particularities in neural activity, autonomic responding, and cognitive risk evaluations. Frontiers in Human Neuroscience, 7: 194. doi: 10.3389/fnhum.2013.00194Google Scholar
Ax, A. (1953). The physiological differentiation between fear and anger in humans. Psychosomatic Medicine, 15: 433442.Google Scholar
Bach, D. R. (2014). Sympathetic nerve activity can be estimated from skin conductance responses: a comment on Henderson et al. (2012). NeuroImage, 84: 122123.Google Scholar
Bechara, A., Damasio, A. R., Damasio, H., & Anderson, S. W. (1994). Insensitivity to future consequences following damage to human prefrontal cortex. Cognition, 50: 715.Google Scholar
Bechara, A., Damasio, H., Damasio, A. R., & Lee, G. P. (1999). Different contributions of the human amygdala and ventromedial prefrontal cortex to decision-making. Journal of Neuroscience, 19: 54735481.Google Scholar
Bechara, A., Damasio, H., Tranel, D., & Damasio, A. R. (1997). Deciding advantageously before knowing the advantageous strategy. Science, 275: 12931295.Google Scholar
Bechara, A., Tranel, D., Damasio, H., Adolphs, R., Rockland, C., & Damasio, A. (1995). Double dissociation of conditioning and declarative knowledge relative to the amygdala and hippocampus in humans. Science, 269: 11151118.Google Scholar
Benedek, M. & Kaernbach, C. (2010). Decomposition of skin conductance data by means of nonnegative deconvolution. Psychophysiology, 47: 647658.Google Scholar
Ben-Shakhar, G. (1985). Standardization within individuals: a simple method to neutralize individual differences in skin conductance. Psychophysiology, 22: 292299.Google Scholar
Bernstein, A. S., Frith, C., Gruzelier, J., Patterson, T., Straube, E., Venables, P., & Zahn, T. (1982). An analysis of the skin conductance orienting response in samples of American, British, and German schizophrenics. Biological Psychology, 14: 155211.Google Scholar
Bernstein, A. S., Taylor, K. W., Starkey, P., Juni, S., Lubowsky, J., & Paley, H. (1981). Bilateral skin conductance, finger pulse volume, and EEG orienting response to tones of differing intensities in chronic schizophrenics and controls. Journal of Nervous and Mental Disease, 169: 513528.Google Scholar
Blair, R. J., Jones, L., Clark, F., & Smith, M. (1997). The psychopathic individual: a lack of responsiveness to distress cues?Psychophysiology, 34: 192198.Google Scholar
Bloch, V. (1993). On the centennial of the discovery of electrodermal activity. In Roy, J. C., Boucsein, W., Fowles, D. C., & Gruzelier, J. H. (eds.), Progress in Electrodermal Research (pp. 16). New York: Plenum Press.Google Scholar
Boucsein, W. (2012). Electrodermal Activity, 2nd edn. New York: Springer.Google Scholar
Boucsein, W., Fowles, D. C., Grimnes, S., Ben-Shakhar, G., Roth, W. T., Dawson, M. E., & Filion, D. L. (2012). Publication recommendations for electrodermal measurements. Psychophysiology, 49: 10171034.Google Scholar
Boyd, R. W. & DiMascio, A. (1954). Social behavior and autonomic physiology: a sociophysiologic study. Journal of Nervous and Mental Disease, 120: 207212.Google Scholar
Bradley, M. M. (2009). Natural selective attention: orienting and emotion. Psychophysiology, 46: 111.Google Scholar
Bradley, M. M. & Lang, P. J. (1999). International Affective Digitized Sounds (IADS): Stimuli, Instruction Manual and Affective Ratings. Technical Report No. B-2, Center for Research in Psychophysiology, University of Florida, Gainesville.Google Scholar
Brekke, J. S., Raine, A., Ansel, M., Lencz, T., & Bird, L. (1997). Neuropsychological and psychophysiological correlates of psychosocial functioning in schizophrenia. Schizophrenia Bulletin, 23: 1928.Google Scholar
Brekke, J. S., Raine, A., & Thomson, C. (1995). Cognitive and psychophysiological correlates of positive, negative, and disorganized symptoms in the schizophrenia spectrum. Psychiatry Research, 57: 241250.Google Scholar
Brown, G., Birley, J. L. T., & Wing, J. K. (1972). Influence of family life on the course of schizophrenia. British Journal of Psychiatry, 121: 241248.Google Scholar
Cheng, D. T., Knight, D. C., Smith, C. N., & Helmstetter, F. J. (2006). Human amygdala activity during the expression of fear responses. Behavioral Neuroscience, 120: 11871195.Google Scholar
Cheng, D. T., Richards, J., & Helmstetter, F. J. (2007). Activity in the human amygdala corresponds to early, rather than late period autonomic responses to a signal for shock. Learning & Memory, 14: 485490.Google 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. 173186). New York: Plenum Press.Google Scholar
Crider, A. (2008). Personality and electrodermal response lability: an interpretation. Applied Psychophysiology and Biofeedback, 33: 141148.Google Scholar
Crider, A. & Augenbraun, C. B. (1975). Auditory vigilance correlates of electrodermal response habituation speed. Psychophysiology, 12: 3640.Google Scholar
Crider, A., Kremen, W. S., Xian, H., Jacobson, K. C., Waterman, B., Eisen, S. A., … & Lyons, M. J. (2004). Stability, consistency, and heritability of electrodermal response lability in middle-aged male twins. Psychophysiology, 41: 501509.Google Scholar
Critchley, H. D. (2002). Electrodermal responses: what happens in the brain. Neuroscientist, 8: 132142.Google Scholar
Critchley, H. D. (2009). Psychophysiology of neural, cognitive and affective integration: fMRI and autonomic indicants. International Journal of Psychophysiology, 73: 8894.Google Scholar
Critchley, H. D., Elliot, R., Mathias, C. J., & Dolan, R. J. (2000). Neural activity relating to generation and representation of galvanic skin conductance responses: a functional magnetic resonance imaging study. Journal of Neuroscience, 20: 30333040.Google Scholar
Critchley, H. D., Mathias, C. J., & Dolan, R. J. (2001). Neural activity in the human brain relating to uncertainty and arousal during anticipation. Neuron, 29: 537545.Google Scholar
Cuthbert, B. N., Bradley, M. M., & Lang, P. J. (1996). Probing picture perception: activation and emotion. Psychophysiology, 33: 103111.Google Scholar
Damasio, A. R. (1994). Descartes’ Error: Emotion, Reason, and the Human Brain. New York: Grosset/Putnam.Google Scholar
Darrow, C. W. (1927). Sensory, secretory, and electrical changes in the skin following bodily excitation. Journal of Experimental Psychology, 10: 197226.Google Scholar
Darrow, C. W. (1937). Neural mechanisms controlling the palmar galvanic skin reflex and palmar sweating. Archives of Neurology and Psychiatry, 37: 641663.Google Scholar
Davies, D. R. & Parasuraman, R. (1982). The Psychology of Vigilance. London: Academic Press.Google Scholar
Dawson, M. E. & Biferno, M. A. (1973). Concurrent measurement of awareness and electrodermal classical conditioning. Journal of Experimental Psychology, 101: 5562.Google Scholar
Dawson, M. E., Gitlin, M., Schell, A. M., Nuechterlein, K. H., & Ventura, J. (1994). Autonomic abnormalities in schizophrenia: state or trait indicators? Archives of General Psychiatry, 51: 813824.Google Scholar
Dawson, M. E. & Nuechterlein, K. H. (1984). Psychophysiological dysfunctions in the developmental course of schizophrenic disorders. Schizophrenia Bulletin, 10: 204232.Google Scholar
Dawson, M. E., Nuechterlein, K. H., & Schell, A. M. (1992a). Electrodermal anomalies in recent-onset schizophrenia: relationships to symptoms and prognosis. Schizophrenia Bulletin, 18: 295311.Google Scholar
Dawson, M. E., Neuchterlein, K. H., Schell, A. M., & Mintz, J. (1992b). Concurrent and predictive electrodermal correlates of symptomatology in recent-onset schizophrenic patients. Journal of Abnormal Psychology, 101: 153164.Google Scholar
Dawson, M. E. & Schell, A. M. (1985). Information processing and human autonomic classical conditioning. In Ackles, P. K., Jennings, J. R., & Coles, M. G. H. (eds.), Advances in Psychophysiology, Volume 1 (pp. 89165). Greenwich, CT: JAI Press.Google Scholar
Dawson, M. E. & Schell, A. M. (2002). What does electrodermal activity tell us about prognosis in the schizophrenia spectrum?Schizophrenia Research, 54: 8793.Google Scholar
Dawson, M. E., Schell, A. M., Rissling, A., Ventura, J., Subotnik, K. L., & Nuechterlein, K. H. (2010). Psychophysiological prodromal signs of schizophrenic relapse: a pilot study. Schizophrenia Research, 123: 6467.Google Scholar
Dethier, V., Bruneau, N., & Philippot, P. (2015). Attentional focus during exposure in spider phobia: the role of schematic versus non-schematic imagery. Behaviour Research and Therapy, 65: 8692.Google Scholar
Dittes, J. E. (1957). Galvanic skin response as a measure of patient’s reaction to therapist’s permissiveness. Journal of Abnormal and Social Psychology, 55: 295303.Google Scholar
Dubé, A., Duquette, M., Roy, M., Lepore, F., Duncan, G., & Rainville, P. (2009). Brain activity associated with the electrodermal reactivity to acute heat pain. NeuroImage, 45: 169180.Google Scholar
Edelberg, R. (1967). Electrical properties of the skin. In Brown, C. C. (ed.), Methods in Psychophysiology (pp. 153). Baltimore, MD: Williams & Wilkens.Google Scholar
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. 367418). New York: Holt.Google Scholar
Edelberg, R. (1993). Electrodermal mechanisms: a critique of the two-effector hypothesis and a proposed replacement. In Roy, J. C., Boucsein, W., Fowles, D. C., & Gruzelier, J. H. (eds.), Progress in Electrodermal Research (pp. 729). New York: Plenum Press.Google Scholar
Esteves, F., Parra, C., Dimberg, U., & Ohman, A. (1994). Nonconscious associative learning: Pavlovian conditioning of skin conductance responses to masked fear-relevant facial stimuli. Psychophysiology, 31: 375385.Google Scholar
Fan, J., Xu, P., Van Dam, N. T., Eilam-Stock, T., Gu, X., Luo, Y., & Hof, P. R. (2012). Spontaneous brain activity relates to autonomic arousal. Journal of Neuroscience, 32: 1117611186.Google Scholar
Féré, C. (1888). Note on changes in electrical resistance under the effect of sensory stimulation and emotion. Comptes Rendus des Séances de la Société de Biologie, (Series 9), 5: 217219.Google Scholar
Ferguson, G. A. & Takane, Y. (1989). Statistical Analysis in Psychology and Education, 6th edn. New York: McGraw-Hill.Google Scholar
Fletcher, R. R., Dobson, K., Goodwin, M. S., Eydgahi, H., Wilder-Smith, O., Fernholz, D., … & Picard, R. W. (2010). iCalm: wearable sensor and network architecture for wirelessly communicating and logging autonomic activity. IEEE Transactions on Information Technology in Biomedicine, 11: 215223.Google Scholar
Fowles, D. C. (1974). Mechanisms of electrodermal activity. In Thompson, R. F. & Patterson, M. M. (eds.), Methods in Physiological Psychology, Part C: Receptor and Effector Processes (pp. 231271). New York: Academic Press.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. 5196). New York: Guilford Press.Google Scholar
Fowles, D. C. (1993). Electrodermal activity and antisocial behavior: empirical findings and theoretical issues. In Roy, J. C., Boucsein, W., Fowles, D. C., & Gruzelier, J. H. (eds.), Progress in Electrodermal Research (pp. 223237). New York: Plenum Press.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: 232239.Google Scholar
Freedman, L. W., Scerbo, A. S., Dawson, M. E., Raine, A., McClure, W. O., & Venables, P. H. (1994). The relationship of sweat gland count to electrodermal activity. Psychophysiology, 31: 196200.Google Scholar
Freixa i Baque, E. (1982). Reliability of electrodermal measures: a compilation. Biological Psychology, 14: 219229.Google Scholar
Frith, C. D., Stevens, M., Johnstone, E. C., & Crow, T. J. (1979). Skin conductance responsivity during acute episodes of schizophrenia as a predictor of symptomatic improvement. Psychological Medicine, 9: 101106.Google Scholar
Fuentes, I., Merita, M. G., Miquel, M., & Rojo, J. (1993). Relationships between electrodermal activity and symptomatology in schizophrenia. Psychopathology, 26: 4752.Google Scholar
Fung, M. T., Raine, A., Loeber, R., Lynam, D. R., Steinhauer, S. R., Venables, P. H., & Stouthamer-Loeber, M. (2005). Reduced electrodermal activity in psychopathy-prone adolescents. Journal of Abnormal Psychology, 114: 187196.Google Scholar
Gao, Y., Raine, A., Venables, P. H., Dawson, M. E., & Mednick, S. A. (2010). Association of poor childhood fear conditioning and adult crime. American Journal of Psychiatry, 167: 5660.Google Scholar
Gentil, A. F., Eskandar, E. N., Marci, C. D., Evans, K. C., & Dougherty, D. D. (2009). Physiological responses to brain stimulation during limbic surgery: further evidence of anterior cingulate modulation of autonomic arousal. Biological Psychiatry, 66: 695701.Google Scholar
Green, M. F., Nuechterlein, K. H., & Satz, P. (1989). The relationship of symptomatology and medication to electrodermal activity in schizophrenia. Psychophysiology, 26: 148157.Google Scholar
Green, S. R., Kragel, P. A., Fecteau, M. E., & LaBar, K. S. (2014). Development and validation of an unsupervised scoring system (autonomate) for skin conductance response analysis. International Journal of Psychophysiology, 91: 186193.Google Scholar
Grey, S. J. & Smith, B. L. (1984). A comparison between commercially available electrode gels and purpose-made gel, in the measurement of electrodermal activity. Psychophysiology, 21: 551557.Google Scholar
Grings, W. W. (1974). Recording of electrodermal phenomena. In Thompson, R. F. & Patterson, M. M. (eds.), Bioelectric Recording Technique, Part C: Receptor and Effector Processes (pp. 273296). New York: Academic Press.Google Scholar
Grings, W. W. & Lockhart, R. A. (1965). Problems of magnitude measurement with multiple GSRs. Psychological Reports, 17: 979982.Google Scholar
Grings, W. W. & Schell, A. M. (1969). Magnitude of electrodermal response to a standard stimulus as a function of intensity and proximity of a prior stimulus. Journal of Comparative and Physiological Psychology, 67: 7782.Google Scholar
Gross, J. J. (1998). Antecedent- and response-focused emotion regulation: divergent consequences for experience, expression, and physiology. Journal of Personality and Social Psychology, 74: 224237.Google Scholar
Gross, J. J. & Levenson, R. W. (1993). Emotional suppression: physiology, self-report, and expressive behavior. Journal of Personality and Social Psychology, 64: 970986.Google Scholar
Haddad, A. D. M., Pritchett, D., Lissek, S., & Lau, J. Y. F. (2012). Trait anxiety and fear responses to safety cues: stimulus generalization or sensitization? Journal of Psychopathology and Behavioral Assessment, 34: 323331.Google Scholar
Hare, R. D. (1965). Temporal gradient of fear arousal in psychopaths. Journal of Abnormal Psychology, 70: 442445.Google Scholar
Hare, R. D. (1991). The Hare Psychopathy Checklist – Revised. Toronto: Multi-Health Systems.Google Scholar
Hassett, J. (1978). A Primer of Psychophysiology. San Francisco: W. H. Freeman and Company.Google Scholar
Hastrup, J. L. (1979). Effects of electrodermal lability and introversion on vigilance decrement. Psychophysiology, 16: 302310.Google Scholar
Hazlett, H., Dawson, M. E., Schell, A. M., & Nuechterlein, K. H. (1997). Electrodermal activity as a prodromal sign in schizophrenia. Biological Psychiatry, 41: 111113.Google Scholar
Heeren, A., Reese, H. E., McNally, R. J., & Philippot, P. (2012). Attention training toward and away from threat in social phobia: effects on subjective, behavioral, and physiological measures of anxiety. Behaviour Research and Therapy, 50: 3039.Google Scholar
Hollister, J. M., Mednick, S. A., Brennan, P., & Cannon, T. D. (1994). Impaired autonomic nervous system habituation in those at genetic risk for schizophrenia. Archives of General Psychiatry, 51: 552558.Google Scholar
Hot, P., Naveteur, J., Leconte, P., & Sequeira, H. (1999). Diurnal variations of tonic electrodermal activity. International Journal of Psychophysiology, 33: 223230.Google Scholar
Hugdahl, K. (1984). Hemispheric asymmetry and bilateral electrodermal recordings: a review of the evidence. Psychophysiology, 21: 371393.Google Scholar
Hugdahl, K. (1995). Psychophysiology: The Mind–Body Perspective. Cambridge, MA: Harvard University Press.Google Scholar
Hugdahl, K. & Öhman, A. (1977). Effects of instruction on acquisition and extinction of electrodermal responses to fear-relevant stimuli. Journal of Experimental Psychology: Human Learning and Memory, 3: 608618.Google Scholar
Hultman, C. M., Öhman, A., Öhlund, L. S., Wieselgren, I., Lindström, L. H., & Öst, L. (1996). Electrodermal activity and social network as predictors of outcome of episodes in schizophrenia. Journal of Abnormal Psychology, 105: 626636.Google Scholar
Iacono, W. G., Ficken, J. W., & Beiser, M. (1993). Electrodermal nonresponding in first-episode psychosis as a function of stimulus significance. In Roy, J. C., Boucsein, W., Fowles, D. C., & Gruzelier, J. H. (eds.), Progress in Electrodermal Activity (pp. 239256). New York: Plenum Press.Google Scholar
Iacono, W. G., Ficken, J. W., & Beiser, M. (1999). Electrodermal activation in first-episode psychotic patients and their first-degree relatives. Psychiatry Research, 88: 2539.Google Scholar
Isen, J. D., Iacono, W. G., Malone, S. M., & McGue, M. (2012). Examining electrodermal hyporeactivity as a marker of externalizing psychopathology: a twin study. Psychophysiology, 49: 10391048.Google Scholar
Isen, J. D., Raine, A., Baker, L., Dawson, M., Bezdjian, S., & Lozano, D. I. (2010). Sex-specific association between psychopathic traits and electrodermal reactivity in children. Journal of Abnormal Psychology, 119: 216225.Google Scholar
Ivory, J. D. & Kalyanaraman, S. (2007). The effects of technological advancement and violent content in video games on players’ feelings of presence, involvement, physiological arousal, and aggression. Journal of Communication, 57: 532555.Google Scholar
Jennings, J. R. (1986). Bodily changes during attending. In Coles, M. G. H., Donchin, E., & Porges, S. W. (eds.), Psychophysiology: Systems, Processes, and Applications (pp. 268289). New York: Guilford Press.Google Scholar
Kappeler-Setz, C., Gravenhorst, F., Schumm, J., Arnrich, B., & Tröster, G. (2013). Towards long term monitoring of electrodermal activity in daily life. Personal and Ubiquitous Computing, 17: 261271.Google Scholar
Katkin, E. S. (1975). Electrodermal lability: a psychophysiological analysis of individual differences in response to stress. In Sarason, I. G. & Spielberger, C. D. (eds.), Stress and Anxiety, Volume 2 (pp. 141176). Washington, DC: Aldine.Google Scholar
Katsanis, J. & Iacono, W. G. (1994). Electrodermal activity and clinical status in chronic schizophrenia. Journal of Abnormal Psychology, 103: 777783.Google Scholar
Kelsey, R. M. (1991). Electrodermal lability and myocardial reactivity to stress. Psychophysiology, 28: 619631.Google Scholar
Kim, D. K., Shin, Y. M., Kim, C. E., Cho, H. S., & Kim, Y. S. (1993). Electrodermal responsiveness, clinical variables, and brain imaging in male chronic schizophrenics. Biological Psychiatry, 33: 786793.Google Scholar
Kivikangas, J. M., Chanel, G., Cowley, B., Ekman, I., Salminen, M., Järvelä, S., & Ravaja, N. (2011). A review of the use of psychophysiological methods in game research. Journal of Gaming and Virtual Worlds, 3: 181199.Google Scholar
Koelega, H. S. (1990). Vigilance performance: a review of electrodermal predictors. Perceptual and Motor Skills, 70: 10111029.Google Scholar
Kring, A. M. & Elis, O. (2013). Emotion deficits in people with schizophrenia. Annual Review of Clinical Psychology, 9: 409433.Google Scholar
Kring, A. M. & Neale, J. M. (1996). Do schizophrenic patients show a disjunctive relationship among expressive, experiential, and psychophysiological components of emotion? Journal of Abnormal Psychology, 105: 249257.Google Scholar
Krzywicki, A. T., Berntson, G. G., & O’Kane, B. L. (2014). A non-contact technique for measuring eccrine sweat gland activity using passive thermal imaging. International Journal of Psychophysiology, 94: 2534.Google Scholar
LaBar, K. S., Gatenby, J. C., Gore, J. C., LeDoux, J. E., & Phelps, E. A. (1998). Human amygdala activation during conditioned fear acquisition and extinction: a mixed-trial fMRI study. Neuron, 20: 937945.Google Scholar
Lacey, J. I., Kagan, J., Lacey, B. C., & Moss, H. A. (1963). The visceral level: situational determinants and behavioral correlates of autonomic response patterns. In Knapp, P. H. (ed.), Expression of the Emotions in Man (pp. 161196). New York: International Universities Press.Google Scholar
Lacey, J. I. & Lacey, B. C. (1958). Verification and extension of the principle of autonomic response-stereotypy. American Journal of Psychology, 71: 5073.Google Scholar
Lader, M. H. & Wing, L. (1966). Psychological Measures, Sedative Drugs, and Morbid Anxiety. London: Oxford University Press.Google Scholar
Landis, C. (1930). Psychology of the psychogalvanic reflex. Psychological Review, 37: 381398.Google Scholar
Lang, P. J., Bradley, M. M., & Cuthbert, B. N. (1998). International Affective Picture System (IAPS): Technical Manual and Affective Ratings. Center for Research in Psychophysiology, University of Florida, Gainesville.Google Scholar
Lang, P. J., Greenwald, M. K., Bradley, M. M., & Hamm, A. O. (1993). Looking at pictures: affective, facial, visceral, and behavioral reactions. Psychophysiology, 30: 261273.Google Scholar
Levenson, R. W. & Gottman, J. M. (1983). Marital interaction: physiological linkage and affective exchange. Journal of Personality and Social Psychology, 45: 587597.Google Scholar
Levenson, R. W. & Gottman, J. M. (1985). Physiological and affective predictors of change in relationship satisfaction. Journal of Personality and Social Psychology, 49: 8594.Google Scholar
Lim, S. & Reeves, B. (2009). Being in the game: effects of avatar choice and point of view on psychophysiological responses during play. Media Psychology, 12: 348370.Google Scholar
Lim, S. & Reeves, B. (2010). Computer agents versus avatars: responses to interactive game characters controlled by a computer or other player. International Journal of Human–Computer Studies, 68: 5768.Google Scholar
Lim, C. L., Rennie, C., Barry, R. J., Bahramali, H., Lazzaro, I., Manor, B., & Gordon, E. (1997). Decomposing skin conductance into tonic and phasic components. International Journal of Psychophysiology, 25: 97109.Google Scholar
Lissek, S., Powers, A. S., McClure, E. B., Phelps, E. A., Woldehawariat, G., Grillon, C., & Pine, D. S. (2005). Classical fear conditioning in the anxiety disorders: a meta-analysis. Behaviour Research and Therapy, 43: 13911424.Google Scholar
Lorber, M. F. (2004). Psychophysiology of aggression, psychopathy, and conduct problems: a meta-analysis. Psychological Bulletin, 130: 531552.Google Scholar
Lovibond, P. F. & Shanks, D. R. (2002). The role of awareness in Pavlovian conditioning: empirical evidence and theoretical implications. Journal of Experimental Psychology: Animal Behavior Processes, 28: 326.Google Scholar
Lykken, D. T. (1957). A study of anxiety in the sociopathic personality. Journal of Abnormal and Social Psychology, 55: 610.Google Scholar
Lykken, D. T. (1959). The GSR in the detection of guilt. Journal of Applied Psychology, 43: 383388.Google Scholar
Lykken, D. T. (1974). Psychology and the lie detector industry. American Psychologist, 29: 725739.Google Scholar
Lykken, D. T. (1995). The Antisocial Personalities. Hillsdale, NJ: Lawrence Erlbaum Associates.Google Scholar
Lykken, D. T., Rose, R. J., Luther, B., & Maley, M. (1966). Correcting psychophysiological measures for individual differences in range. Psychological Bulletin, 66: 481484.Google Scholar
Lykken, D. T. & Venables, P. H. (1971). Direct measurement of skin conductance: a proposal for standardization. Psychophysiology, 8: 656672.Google Scholar
Lyytinen, H., Blomberg, A. P., & Näätänen, R. (1992). Event-related potentials and autonomic responses to a change in unattended auditory stimuli. Psychophysiology, 29: 523534.Google Scholar
Mangina, C. A. & Beuzeron-Mangina, J. H. (1996). Direct electrical stimulation of specific human brain structures and bilateral electrodermal activity. International Journal of Psychophysiology, 22: 18.Google Scholar
Matthews, A., Naran, N., & Kirkby, K. C. (2015). Symbolic online exposure for spider fear: habituation of fear, disgust and physiological arousal and predictors of symptom improvement. Journal of Behavior Therapy and Experimental Psychiatry, 47: 129137.Google Scholar
Meijer, E. H., Selle, N. K., Elber, L., & Ben-Shakhar, G. (2014). Memory detection with the concealed information test: a meta analysis of skin conductance, respiration, heart rate, and P300 data. Psychophysiology, 51: 879904.Google Scholar
Montague, J. D. (1963). Habituation of the psycho-galvanic reflex during serial tests. Journal of Psychosomatic Research, 7: 199214.Google Scholar
Morris, J. S., Buchel, C., & Dolan, R. J. (2001). Parallel neural responses in amygdala subregions and sensory cortex during implicit fear conditioning. NeuroImage, 13: 10441052.Google Scholar
Moscovitch, D. A., Suvak, M. K., & Hofmann, S. G. (2010). Emotional response patterns during social threat in individuals with generalized social anxiety disorder and non-anxious controls. Journal of Anxiety Disorders, 24: 785791.Google Scholar
Mosig, C., Merz, C. J., Mohr, C., Adolph, D., Wolf, O. T., Schneider, S., & Zlomuzica, A. (2014). Enhanced discriminative fear learning of phobia-irrelevant stimuli in spider-fearful individuals. Frontiers in Behavioral Neuroscience, 8: 328. doi:10.3389/fnbeh.2014.00328Google Scholar
Mundy-Castle, A. C. & McKiever, B. L. (1953). The psychophysiological significance of the galvanic skin response. Journal of Experimental Psychology, 46: 1524.Google Scholar
Munro, L. L., Dawson, M. E., Schell, A. M., & Sakai, L. M. (1987). Electrodermal lability and rapid performance decrement in a degraded stimulus continuous performance task. Journal of Psychophysiology, 1: 249257.Google Scholar
Neumann, E. & Blanton, R. (1970). The early history of electrodermal research. Psychophysiology, 6: 453475.Google Scholar
Nuechterlein, K. H. & Dawson, M. E. (1984). A heuristic vulnerability/stress model of schizophrenic episodes. Schizophrenia Bulletin, 10: 300312.Google Scholar
Öhman, A. (1981). Electrodermal activity and vulnerability to schizophrenia: a review. Biological Psychology, 12: 87145.Google Scholar
Öhman, A. (2009). Of snakes and faces: an evolutionary perspective on the psychology of fear. Scandinavian Journal of Psychology, 50: 543552.Google Scholar
Öhman, A., Dimberg, U., & Esteves, F. (1989a). Preattentive activation of aversive emotions. In Archer, T. & Nilsson, L. G. (eds.), Aversion, Avoidance, and Anxiety (pp. 169193). Hillsdale, NJ: Lawrence Erlbaum Associates.Google Scholar
Öhman, A. & Mineka, S. (2001). Fears, phobias, and preparedness: toward an evolved module of fear and fear learning. Psychological Review, 108: 483522.Google Scholar
Öhman, A., Öhlund, L. S., Alm, T., Wieselgren, I., Öst, L., & Lindström, L. H. (1989b). Electrodermal nonresponding, premorbid adjustment, and symptomatology as predictors of long-term social functioning in schizophrenics. Journal of Abnormal Psychology, 98: 426435.Google Scholar
Öhman, A. & Soares, J. J. F. (1998). Emotional conditioning to masked stimuli: expectancies for aversive outcomes following nonrecognized fear-relevant stimuli. Journal of Experimental Psychology: General, 127: 6982.Google Scholar
Patterson, J. C., Ungerleider, L. G., & Bandettini, P. A. (2002). Task-independent functional brain activity correlation with skin conductance changes: an fMRI study. NeuroImage, 17: 17971806.Google Scholar
Payne, A. F. H., Dawson, M. E., Schell, A. M., Singh, K., & Courtney, C. G. (2013). Can you give me a hand? A comparison of hands and feet as optimal anatomical sites for skin conductance recording. Psychophysiology, 50: 10651069.Google Scholar
Payne, A. F., Schell, A. M., & Dawson, M. E. (2016). Lapses in skin conductance responding across anatomical sites: comparison of fingers, feet, forehead, and wrist. Psychophysiology, 53: 10841092.Google Scholar
Phelps, E. A., Delgado, M. R., Nearing, K. I., & LeDoux, J. E. (2004). Extinction learning in humans: role of the amygdala and vmPFC. Neuron, 43: 897905.Google Scholar
Picard, R. W., Fedor, S., & Ayzenberg, Y. (2015). Multiple arousal theory and daily-life electrodermal activity asymmetry. Emotion Review, 8. doi: 10.1177/1754073914565517Google Scholar
Picard, R. W. & Healey, J. (1997). Affective wearables. Personal Technologies, 1: 231240.Google Scholar
Poh, M., Swenson, N. C., & Picard, R. W. (2010). A wearable sensor for unobtrusive, long-term assessment of electrodermal activity. IEEE Transactions on Biomedical Engineering, 57: 12431252.Google Scholar
Pole, N. (2007). The psychophysiology of posttraumatic stress disorder: a meta-analysis. Psychological Bulletin, 133: 725746.Google Scholar
Prokasy, W. F. & Kumpfer, K. L. (1973). Classical conditioning. In Prokasy, W. F. & Raskin, D. C. (eds.), Electrodermal Activity in Psychological Research (pp. 157202). New York: Academic Press.Google Scholar
Prokasy, W. F. & Raskin, D. C. (eds.) (1973). Electrodermal Activity in Psychological Research. New York: Academic.Google Scholar
Quay, H. C. (1965). Psychopathic personality as pathological stimulation-seeking. American Journal of Psychiatry, 122: 180183.Google Scholar
Raby, K. L., Roisman, G. I., Simpson, J. A., Collins, W. A., & Steele, R. D. (2015). Greater maternal insensitivity in childhood predicts greater electrodermal reactivity during conflict discussions with romantic partners in adulthood. Psychological Science, 26: 348353.Google Scholar
Raine, A., Venables, P. H., & Williams, M. (1990). Relationships between central and autonomic measures of arousal at age 15 years and criminality at age 24 years. Archives of General Psychiatry, 47: 10031007.Google Scholar
Roisman, G. I. (2007). The psychophysiology of adult attachment relationships: autonomic reactivity in marital and premarital interactions. Developmental Psychology, 43: 3953.Google Scholar
Rothemund, Y., Ziegler, S., Hermann, C., Gruesser, S. M., Foell, J., Patrick, C. J., & Flor, H. (2012). Fear conditioning in psychopaths: event-related potentials and peripheral measures. Biological Psychology, 90: 5059.Google Scholar
Roy, J. C., Sequeira, H., & Delerm, B. (1993). Neural control of electrodermal activity: spinal and reticular mechanisms. In Roy, J. C., Boucsein, W., Fowles, D. C., & Gruzelier, J. H. (eds.), Progress in Electrodermal Research (pp. 7392). New York: Plenum Press.Google Scholar
Sakai, L. M., Baker, L. A., & Dawson, M. E. (1992). Electrodermal lability: individual differences affecting perceptual speed and vigilance performance in 9 to 16 year-old children. Psychophysiology, 29: 207217.Google Scholar
Sano, A., Picard, R. W., & Stickgold, R. (2014). Quantitative analysis of wrist electrodermal activity during sleep. International Journal of Psychophysiology, 94: 382389.Google Scholar
Scerbo, A. S., 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: 241246.Google Scholar
Schell, A. M., Dawson, M. E., & Filion, D. L. (1988). Psychophysiological correlates of electrodermal lability. Psychophysiology, 25: 619632.Google Scholar
Schell, A. M., Dawson, M. E., & Marinkovic, K. (1991). Effects of potentially phobic conditioned stimuli on retention, reconditioning, and extinction of the conditioned skin conductance response. Psychophysiology, 28: 140153.Google Scholar
Schell, A. M., Dawson, M. E., Nuechterlein, K. H., Subotnik, K. L., & Ventura, J. (2002). The temporal stability of electrodermal variables over a one-year period in patients with recent-onset schizophrenia and in normal subjects. Psychophysiology, 39: 124132.Google Scholar
Schell, A. M., Dawson, M. E., Rissling, A., Ventura, J., Subotnik, K. L., Gitlin, M. J., & Nuechterlein, K. H. (2005). Electrodermal predictors of functional outcome and negative symptoms in schizophrenia. Psychophysiology, 42: 483492.Google Scholar
Schumm, J., Bachlin, M., Setz, C., Arnrich, B., Roggen, D., & Troster, G. (2008). Effect of movements on the electrodermal response after a startle event. Methods of Information in Medicine, 47: 186191.Google Scholar
Seligman, M. E. (1970). On the generality of the laws of learning. Psychological Review, 77: 406418.Google Scholar
Sequeira, H. & Roy, J.-C. (1993). Cortical and hypothalamo-limbic control of electrodermal responses. In Roy, J. C., Boucsein, W., Fowles, D. C., & Gruzelier, J. H. (eds.), Progress in Electrodermal Research (pp. 93114). New York: Plenum Press.Google Scholar
Shiban, Y., Pauli, P., & Mühlberger, A. (2013). Effect of multiple context exposure on renewal in spider phobia. Behaviour Research and Therapy, 51: 6874.Google Scholar
Shields, S. A., MacDowell, K. A., Fairchild, S. B., & Campbell, M. L. (1987). Is mediation of sweating cholinergic, adrenergic, or both? A comment on the literature. Psychophysiology, 24: 312319.Google Scholar
Siddle, D., Stephenson, D., & Spinks, J. A. (1983). Elicitation and habituation of the orienting response. In Siddle, D. (ed.), Orienting and Habituation: Perspectives in Human Research (pp. 109182). Chichester: John Wiley.Google Scholar
Sokolov, E. N. (1963). Perception and the Conditioned Reflex. New York: Macmillan.Google Scholar
Spohn, H. E., Coyne, L., Wilson, J. K., & Hayes, K. (1989). Skin-conductance orienting response in chronic schizophrenics: the role of neuroleptics. Journal of Abnormal Psychology, 98: 478486.Google Scholar
Stern, R. M., Ray, W. J., & Quigley, K. S. (2001). Psychophysiological Recording. Oxford University Press.Google Scholar
Straube, E. R. (1979). On the meaning of electrodermal nonresponding in schizophrenia. Journal of Nervous and Mental Disease, 167: 601611.Google Scholar
Subotnik, K. I., Schell, A. M., Chilingar, M. S., Dawson, M. E., Ventura, J., Kelly, K. A., … & Nuechterlein, K. H. (2012). The interaction of electrodermal activity and expressed emotion in predicting symptoms in recent-onset schizophrenia. Psychophysiology, 49: 10351038.Google Scholar
Tarchanoff, J. (1890). Galvanic phenomena in the human skin during stimulation of the sensory organs and during various forms of mental activity. Pflugers Archive für die Gesamte Physiologie des Menschen und der Tiere, 46: 4655.Google Scholar
Tarrier, N. & Barrowclough, C. (1989). Electrodermal activity as a predictor of schizophrenic relapse. Psychopathology, 22: 320324.Google Scholar
Tarrier, N., Vaughn, C., Lader, M. H., & Leff, J. P. (1979). Bodily reactions to people and events in schizophrenics. Archives of General Psychiatry, 36: 311315.Google Scholar
Tartz, R., Vartak, A., King, J., & Fowles, D. (2015). Effects of grip force on skin conductance measured from a handheld device. Psychophysiology, 52: 819.Google Scholar
Tranel, D. (2000). Electrodermal activity in cognitive neuroscience: neuroanatomical and neurophysiological correlates. In Lane, R. D. & Nadel, L. (eds.), Cognitive Neuroscience of Emotion (pp. 192224). Oxford University Press.Google Scholar
Vaidyanathan, U., Isen, J. D., Malone, S. M., Miller, M. B., McGue, M., & Iacono, W. G. (2014). Heritability and molecular genetic basis of electrodermal activity: a genome-wide association study. Psychophysiology, 51: 12591271.Google Scholar
Van Bockstaele, B., Verschuere, B., Koster, E. H. W., Tibboel, H., De Houwer, J., & Crombez, G. (2011). Effects of attention training on self-reported, implicit, physiological and behavioural measures of spider fear. Journal of Behavior Therapy and Experimental Psychiatry, 42: 211218.Google Scholar
van Dooren, M., de Vries, J. J., & Janssen, J. H. (2012). Emotional sweating across the body: comparing 16 different skin conductance measurement locations. Physiology & Behavior, 106: 298304.Google Scholar
Vaughn, C. E. & Leff, J. (1976). The influence of family and social factors on the course of psychiatric illness: a comparison of schizophrenic and depressed neurotic patients. British Journal of Psychiatry, 129: 125137.Google Scholar
Vaughn, C. E., Snyder, K. S., Jones, S., Freeman, W. B., & Falloon, I. R. H. (1984). Family factors in schizophrenic relapse: a California replication of the British research on expressed emotion. Archives of General Psychiatry, 41: 11691177.Google Scholar
Veit, R., Konica, L., Klinzing, J. G., Barth, B., Yilmaz, O., & Birbaumer, N. (2013). Deficient fear conditioning in psychopathy as a function of interpersonal and affective disturbances. Frontiers in Human Neuroscience, 7: 706. doi:10.3389/fnhum.2013.00706Google Scholar
Venables, P. H. & Christie, M. J. (1973). Mechanisms, instrumentation, recording techniques, and quantification of responses. In Prokasy, W. F. & Raskin, D. C. (eds.), Electrodermal Activity in Psychological Research (pp. 1124). New York: Academic Press.Google Scholar
Venables, P. H. & Christie, M. J. (1980). Electrodermal activity. In Martin, I. & Venables, P. H. (eds.), Techniques in Psychophysiology (pp. 367). Chichester: John Wiley.Google Scholar
Venables, P. H. & Mitchell, D. A. (1996). The effects of age, sex and time of testing on skin conductance activity. Biological Psychology, 43: 87101.Google Scholar
Verona, E., Patrick, C. J., Curtin, J., Bradley, M. M., & Lang, P. J. (2004). Psychopathy and physiological responses to emotionally evocative sounds. Journal of Abnormal Psychology, 113: 99108.Google Scholar
Verschuere, B. & Ben-Shakhar, G. (2011). Theory of the concealed information test. In Verschuere, B., Ben-Shakhar, G., & Meijer, E. H. (eds.), Theory and Application of the Concealed Information Test (pp. 128148). Cambridge University Press.Google Scholar
Verschuere, B., Ben-Shakhar, G., & Meijer, E. H. (eds.). (2011). Memory Detection: Theory and Application of the Concealed Information Test. Cambridge University Press.Google Scholar
Vigouroux, R. (1879). Sur le rôle de la résistance électrique des tissues dans l’electro-diagnostic. Comptes Rendus Société de Biologie, 31: 336339.Google Scholar
Vigouroux, R. (1888). The electrical resistance considered as a clinical sign. Progrès Medicale, 3: 8789.Google Scholar
Vossel, G. & Rossman, R. (1984). Electrodermal habituation speed and visual monitoring performance. Psychophysiology, 21: 97100.Google Scholar
Wallin, B. G. (1981). Sympathetic nerve activity underlying electrodermal and cardiovascular reactions in man. Psychophysiology, 18: 470476.Google Scholar
Wang, G. H. (1964). The Neural Control of Sweating. Madison, WI: University of Wisconsin Press.Google Scholar
Wieselgren, I. M., Öhlund, L. S., Lindstrom, L. H., & Öhman, A. (1994). Electrodermal activity as a predictor of social functioning in female schizophrenics. Journal of Abnormal Psychology, 103: 570573.Google Scholar
Wilhelm, F. H., Pfaltz, M. C., Grossman, P., & Roth, W. T. (2006). Distinguishing emotional from physical activation in ambulatory psychophysiological monitoring. Behavioral Sciences Instrumentation, 42: 458463.Google Scholar
Wilhelm, F. H. & Roth, W. T. (1998). Taking the laboratory to the skies: ambulatory assessment of self-report, autonomic, and respiratory responses in flying phobia. Psychophysiology, 35: 596606.Google Scholar
Woodworth, R. S. & Schlosberg, H. (1954). Experimental Psychology, rev. edn. New York: Holt & Co.Google Scholar
Zahn, T. P., Carpenter, W. T., & McGlashan, T. H. (1981). Autonomic nervous system activity in acute schizophrenia: II. Relationships to short-term prognosis and clinical state. Archives of General Psychiatry, 38: 260266.Google Scholar
Zahn, T. P., Grafman, J., & Tranel, D. (1999). Frontal lobe lesions and electrodermal activity: effects of significance. Neuropsychologia, 37: 12271241.Google Scholar
Zhang, S., Hu, S., Chao, H. H., Luo, X., Farr, O. M., & Li, C. S. (2012). Cerebral correlates of skin conductance responses in a cognitive task. NeuroImage, 62: 14891498.Google Scholar

References

Albéri, L. (2013). Asthma: a clinical condition for brain health. Experimental Neurology, 248: 338342.Google Scholar
Alpher, V. S., Nelson, R. B., & Blanton, R. L. (1986). Effects of cognitive and psychomotor tasks on breath-holding span. Journal of Applied Physiology, 61: 11491152.Google Scholar
Arzi, A., Shedlesky, L., Secundo, L., & Sobel, N. (2014). Mirror sniffing: humans mimic olfactory sampling behavior. Chemical Senses, 39: 277281.Google Scholar
Begnignus, V. & Prah, J. D. (1980) A computer-controlled vapor-dilution olfactometer. Behavior Research Methods & Instrumentation, 12: 535540.Google Scholar
Berntson, G. G., Cacioppo, J. T., & Grossman, P. (2007). Whither vagal tone. Biological Psychology, 74: 295300.Google Scholar
Birn, R. M., Smith, M. A., Jones, T. B., & Bandettini, P. A. (2008). The respiration response function: the temporal dynamics of fMRI signal fluctuations related to changes in respiration. NeuroImage, 40: 644654.Google Scholar
Bloch-Salisbury, E., Harver, A., & Squires, N. K. (1998). Event-related potentials to inspiratory flow-resistive loads in young adults: stimulus magnitude effects. Biological Psychology, 49: 165186.Google Scholar
Boiten, F. A., Frijda, N. H., & Wientjes, C. J. (1994). Emotions and respiratory patterns: review and critical analysis. International Journal of Psychophysiology, 17: 103128.Google Scholar
Braman, S. S. (1995). The regulation of normal lung function. Allergy and Asthma Proceedings, 16: 223226).Google Scholar
Brumbaugh, C. C., Kothuri, R., Marci, C., Siefert, C., & Pfaff, D. D. (2013). Physiological correlates of the Big 5: autonomic responses to video presentations. Applied Psychophysiology and Biofeedback, 38: 293301.Google Scholar
Burtt, H. E. (1921). Further technique for inspiration-expiration ratios. Journal of Experimental Psychology, 4: 106110.Google Scholar
Casali, J. G., Wierwille, W. W., & Cordes, R. E. (1983). Respiration measurement: overview and new instrumentation. Behavior Research Methods & Instruments, 15: 401405.Google Scholar
Chang, Y. C. & Huang, S. L. (2012). The influence of attention levels on psychophysiological responses. International Journal of Psychophysiology, 86: 3947.Google Scholar
Chen, E. & Miller, G. E. (2007). Stress and inflammation in exacerbations of asthma. Brain, Behavior, and Immunity, 21: 993999.Google Scholar
Comroe, J. H. (1974). Mechanical factors in breathing. In Physiology of Respiration, 2nd edn. (pp. 94141). Chicago, IL: Yearbook Medical.Google Scholar
Dalton, P. (2002). Olfaction. In Pashler, H. & Yantis, S. (eds.), Stevens’ Handbook of Experimental Psychology, Volume 1 (pp. 691746). New York: John Wiley.Google Scholar
Elmes, D. G. & Lorig, T. S. (2008). Adequacy of control comparisons in olfactory experiments. Chemosensory Perception, 1: 247252.Google Scholar
Ernst, J. M., Litvack, D. A., Lozano, D. L., Cacioppo, J. T., & Berntson, G. G. (1999). Impedance pneumography: noise as signal in impedance cardiography. Psychophysiology, 36: 333338.Google Scholar
Feleky, A. (1916). The influence of the emotions on respiration. Journal of Experimental Psychology, 1: 218241.Google Scholar
Frank, R. A., Gesteland, R. C., Bailie, J., Rybalsky, K., Seiden, A., & Dulay, M. F. (2006). Characterization of the sniff magnitude test. Archives of Otolaryngology – Head & Neck Surgery, 132: 532536.Google Scholar
Freeman, W. J. & Schneider, W. (1982). Changes in spatial patterns of rabbit olfactory EEG with conditioning to odors. Psychophysiology, 19: 4456.Google Scholar
Freeman, W. J. & Viana Di Prisco, G. (1986). Relation of olfactory EEG to behavior: time series analysis. Behavioral Neuroscience, 100: 753763.Google Scholar
Glower, D. D., Spratt, J. A., Snow, N. D., Kabas, J. S., Davis, J. W., Olsen, C. O., … & Rankin, J. S. (1985). Linearity of the Frank-Starling relationship in the intact heart: the concept of preload recruitable stroke work. Circulation, 71: 9941009.Google Scholar
Gomez, P., Zimmermann, P., Guttormsen-Schär, S., & Danuser, B. (2005). Respiratory responses associated with affective processing of film stimuli. Biological Psychology, 68: 223235.Google Scholar
Gross, R. D., Atwood, C. W. Jr., Ross, S. B., Eichhorn, K. A., Olszewski, J. W., & Doyle, P. J. (2008). The coordination of breathing and swallowing in Parkinson’s disease. Dysphagia, 23: 136145.Google Scholar
Harada, Y., Kuno, M. Y., & Wang, Y. Z. (1985). Differential effects of carbon dioxide and pH on central chemoreceptors in the rat in vitro. Journal of Physiology, 368: 679693.Google Scholar
Harver, A. (1994). Effects of feedback on the ability of asthmatic subjects to detect increases in the flow-resistive component to breathing. Health Psychology, 13: 5262.Google Scholar
Henderson, A., Goldman-Eisler, F., & Skarbek, A. (1965). Temporal patterns of cognitive activity and breath control in speech. Language and Speech, 8: 236242.Google Scholar
Hlastala, M. P. & Berger, A. J. (2001). Chemical control of breathing. In Hlastala, M. P. & Berger, A. J., Physiology of Respiration, 2nd edn. (pp. 148166). Oxford University Press.Google Scholar
Hogg, J. C., Pare, P. D., Boucher, R. C., & Michoud, M. C. (1979). The pathophysiology of asthma. Canadian Medical Association Journal, 121: 409414.Google Scholar
Ismail, R. A. & Babiker, S. F. (2015). Oxygen level measurement techniques: pulse oximetry. Journal of Engineering and Computer Science, 16: 15.Google Scholar
Jamner, L. D., Shapiro, D., Goldstein, I. B., & Hug, R. (1991). Ambulatory blood pressure and heart rate in paramedics: effects of cynical hostility and defensiveness. Psychosomatic Medicine, 53: 393406.Google Scholar
Johnson, R. L. & Miller, J. M. (1968). Distribution of ventilation, blood flow, and gas transfer coefficients in the lung. Journal of Applied Physiology, 25: 115.Google Scholar
Klein, D. F. (1994). Testing the suffocation false alarm theory of panic disorder. Anxiety, 1: 17.Google Scholar
Kobal, G. & Hummel, C. (1988). Cerebral chemosensory evoked potentials elicited by chemical stimulation of the human olfactory and respiratory nasal mucosa. Electroencephalography & Clinical Neurophysiology/Evoked Potentials Section, 71: 241250.Google Scholar
Komisaruk, B. R. (1970). Synchrony between limbic system theta activity and rhythmical behavior in rats. Journal of Comparative and Physiological Psychology, 171: 157192.Google Scholar
Konno, K. & Mead, J. (1967). Measurement of the separate volume changes of rib cage and abdomen during breathing. Journal of Applied Physiology, 22: 407422.Google Scholar
Laing, D. G. (1983). Natural sniffing gives optimum odour perception for humans. Perception, 12: 99117.Google Scholar
Landis, C. & Gullette, R. (1925). Studies of emotional reactions: III. Systolic blood pressure and inspiration-expiration ratios. Journal of Comparative Psychology, 5: 221253.Google Scholar
Li, G., Huang, H., Wei, J., Li, D. G., Chen, Q., Gaebler, C. P., … & Mechalakos, J. (2015). Novel spirometry based on optical surface imaging. Medical Physics, 42: 16901697.Google Scholar
Lorig, T. S., Matia, D. C., Peszka, J. J., & Bryant, D. N. (1996). The effects of active and passive stimulation on chemosensory event-related potentials. International Journal of Psychophysiology, 23: 199205.Google Scholar
Lugaresi, E. & Vela-Bueno, A. (1987). Sleep-related respiratory disorders. Seminars in Neurology, 7: 259268.Google Scholar
Mainland, J. & Sobel, N. (2006). The sniff is part of the olfactory percept. Chemical Senses, 31: 181196.Google Scholar
Masaoka, Y., Satoh, H., Akai, L., & Homma, I. (2010). Expiration: the moment we experience retronasal olfaction in flavor. Neuroscience Letters, 473: 9296.Google Scholar
Murdock, K. K., Robinson, E. M., Adams, S. K., Berz, J., & Rollock, M. J. (2009). Family–school connections and internalizing problems among children living with asthma in urban, low-income neighborhoods. Journal of Child Health Care, 13: 275294.Google Scholar
Muth, E. R., Moss, J. D., Rosopa, P. J., Salley, J. N., & Walker, A. D. (2012). Respiratory sinus arrhythmia as a measure of cognitive workload. International Journal of Psychophysiology, 83: 96101.Google Scholar
Nardi, A. E., Freire, R. C., & Zin, W. A. (2009). Panic disorder and control of breathing. Respiratory Physiology & Neurobiology, 167: 133143.Google Scholar
Nitzan, M., Romem, A., & Koppel, R. (2014). Pulse oximetry: fundamentals and technology update. Medical Devices (Auckland, NZ), 7: 231239.Google Scholar
Pennock, B. E., Cox, C. P., Rogers, R. M., Cain, W. A., & Wells, J. H. (1979). A noninvasive technique for measurement of changes in specific airway resistance. Journal of Applied Physiology, 46: 399406.Google Scholar
Pfaltz, M. C., Michael, T., Grossman, P., Blechert, J., & Wilhelm, F. H. (2009). Respiratory pathophysiology of panic disorder: an ambulatory monitoring study. Psychosomatic Medicine, 71: 869876.Google Scholar
Porges, S. W. (1995). Cardiac vagal tone: a physiological index of stress. Neuroscience & Biobehavioral Reviews, 19: 225233.Google Scholar
Quinn, S. J., Huang, L., Ellis, P. D. M., & Williams, J. F. (1996). The differentiation of snoring mechanisms using sound analysis. Clinical Otolaryngology & Allied Sciences, 21: 119123.Google Scholar
Raichle, M. E. (2010). The brain’s dark energy. Scientific American, 302: 4449.Google Scholar
Richards, D. W. Jr. (1953). The nature of cardiac and pulmonary dyspnea. Circulation, 7: 1529.Google Scholar
Rittweger, J., Lambertz, M., & Langhorst, P. (1997). Influences of mandatory breathing on rhythmical components of electrodermal activity. Clinical Physiology, 17: 609618.Google Scholar
Ritz, T., Dahme, B., Dubois, A. B., Folgering, H., Fritz, G. K., Harver, A., … & Woestijne, K. P. (2002). Guidelines for mechanical lung function measurements in psychophysiology. Psychophysiology, 39: 546567.Google Scholar
Sandberg, S., Paton, J. Y., Ahola, S., McCann, D. C., McGuinness, D., Hillary, C. R., & Oja, H. (2000). The role of acute and chronic stress in asthma attacks in children. The Lancet, 356: 982987.Google Scholar
Sasaki, C. T. & Mann, D. G. (1976). Dilator naris function: a useful test of facial nerve integrity. Archives of Otolaryngology, 102: 365367.Google Scholar
Seoane, F., Mohino-Herranz, I., Ferreira, J., Alvarez, L., Buendia, R., Ayllón, D., … & Gil-Pita, R. (2014). Wearable biomedical measurement systems for assessment of mental stress of combatants in real time. Sensors, 14: 71207141.Google Scholar
Sobel, N., Prabhakaran, V., Hartley, C. A., Desmond, J. E., Zhao, Z., Glover, G. H., … & Sullivan, E. V. (1998). Odorant-induced and sniff-induced activation in the cerebellum of the human. Journal of Neuroscience, 18: 89909001.Google Scholar
Spyer, K. M. (2009). To breathe or not to breathe? That is the question. Experimental Physiology, 94: 110.Google Scholar
Stick, S. M., Ellis, E., LeSouëf, P. N., & Sly, P. D. (1992). Validation of respiratory inductance plethysmography (“Respitrace”®) for the measurement of tidal breathing parameters in newborns. Pediatric Pulmonology, 14: 187191.Google Scholar
Strauss-Blasche, G., Moser, M., Voica, M., McLeod, D. R., Klammer, N., & Marktl, W. (2000). Relative timing of inspiration and expiration affects respiratory sinus arrhythmia. Clinical and Experimental Pharmacology and Physiology, 27: 601606.Google Scholar
Sul, B., Wallqvist, A., Morris, M. J., Reifman, J., & Rakesh, V. (2014). A computational study of the respiratory airflow characteristics in normal and obstructed human airways. Computers in Biology and Medicine, 52: 130143.Google Scholar
Thayer, J. F. & Lane, R. D. (2000). A model of neurovisceral integration in emotion regulation and dysregulation. Journal of Affective Disorders, 61: 201216.Google Scholar
Vanderwolf, C. H. (1992). Hippocampal activity, olfaction, and sniffing: an olfactory input to the dentate gyrus. Brain Research, 593: 197208.Google Scholar
Van Diest, I., Thayer, J. F., Vandeputte, B., Van de Woestijne, K. P., & Van den Bergh, O. (2006). Anxiety and respiratory variability. Physiology & Behavior, 89: 189195.Google Scholar
Verhagen, J. V., Wesson, D. W., Netoff, T. I., White, J. A., & Wachowiak, M. (2007). Sniffing controls an adaptive filter of sensory input to the olfactory bulb. Nature Neuroscience, 10: 631639.Google Scholar
Vlemincx, E., Abelson, J. L., Lehrer, P. M., Davenport, P. W., Van Diest, I., & Van den Bergh, O. (2013). Respiratory variability and sighing: a psychophysiological reset model. Biological Psychology, 93: 2432.Google Scholar
Vlemincx, E., Taelman, J., De Peuter, S., Van Diest, I., & Van Den Bergh, O. (2011). Sigh rate and respiratory variability during mental load and sustained attention. Psychophysiology, 48: 117120.Google Scholar
Wachowiak, M. (2011). All in a sniff: olfaction as a model for active sensing. Neuron, 71: 962973.Google Scholar
Wientjes, C. J. & Grossman, P. (1998). Respiratory psychophysiology as a discipline: introduction to the special issue. Biological Psychology, 49: 18.Google Scholar
Wilhelm, F. H., Roth, W. T., & Sackner, M. A. (2003). The LifeShirt: an advanced system for ambulatory measurement of respiratory and cardiac function. Behavior Modification, 27: 671691.Google Scholar
Woodworth, R. S. & Schlosberg, S. (1965). Experimental Psychology. New York: Holt.Google Scholar
Yeh, H. C. & Schum, G. M. (1980). Models of human lung airways and their application to inhaled particle deposition. Bulletin of Mathematical Biology, 42: 461480.Google Scholar

References

Abell, T. L. & Malagelada, J. R. (1985). Glucagon-evoked gastridysrhythmias in humans shown by an improved electrogastrographic technique. Gastroenterology, 88: 19321940.Google Scholar
Abell, T. L., Malagelada, J. R., Lucas, A. R., Brown, M. L., Camilleri, M., Go, V. L., … & Zinsmeister, A. R. (1987). Gastric electromechanical and neurohormonal function in anorexia nervosa. Gastroenterology, 93: 958965.Google Scholar
Abell, T. L., Tucker, R., & Malagelada, J. R. (1985). Simultaneous gastric electromanometry in man. In Stern, R. M. & Koch, K. L. (eds.), Electrogastrography (pp. 7888). New York: Praeger.Google Scholar
Alvarez, W. C. (1922). The electrogastrogram and what it shows. Journal of the American Medical Association, 78, 11161119.Google Scholar
Alvarez, W. C. (1943). Nervousness, Indigestion, and Pain. New York: Hoeber.Google Scholar
Baldaro, B., Mazzetti, M., Codispoti, M., Tuozzi, G., Bolzani, R., & Trombini, G. (2001). Autonomic reactivity during viewing of an unpleasant film. Perceptual and Motor Skills, 93: 797805.Google Scholar
Beaumont, W. (1959 [1833]). Experiments and Observations on the Gastric Juice and the Physiology of Indigestion. New York: Dover.Google Scholar
Benson, P. W., Hooker, J. B., Koch, K. L., & Weinberg, R. B. (2012). Bitter taster status predicts susceptibility to vection-induced motion sickness and nausea. Neurogastroenterology and Motility, 24: 134140.Google 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: 599608.Google Scholar
Berntson, G. G., Quigley, K. S., & Lozano, D. (2007). Cardiovascular psychophysiology. In Caccioppo, J. T., Tassinary, L. G., & Berntson, G. G. (eds.), Handbook of Psychophysiology, 3rd edn. (pp. 182210). Cambridge University Press.Google Scholar
Bortolotti, M., Sarti, P., Barara, L., & Brunelli, F. (1990). Gastric myoelectrical activity in patients with chronic idiopathic gastroparesis. Journal of Gastrointestinal Motility, 2: 104108.Google Scholar
Brehmer, A. (2006). Structure of enteric neurons. Advances in Anatomy, Embryology, and Cell Biology, 186: 191.Google Scholar
Brown, B. H., Smallwood, R. H., Duthie, H. L., & Stoddard, C. J. (1975). Intestinal smooth muscle electrical potentials recorded from surfaces electrodes. Medical Biological Engineering, 13: 97103.Google Scholar
Bruley des Varannes, S., Mizrahi, M., & Dubois, A. (1991). Relation between postprandial gastric emptying and cutaneous electrogastrogram in primates. American Journal of Physiology, 261: G248G255.Google Scholar
Bruno, C., Lopetuso, L. R., Ianiro, G., Laterza, L., Gerardi, V., Petito, V., … & Scaldaferri, F., (2013). 13C-octanoic acid breath test to study gastric emptying time. European Review for Medical and Pharmacological Sciences, 17: 5964.Google Scholar
Brzana, R. J., Koch, K. L., & Bingaman, S. (1998). Gastric myoelectrical activity in patients with gastric outlet obstruction and idiopathic gastroparesis. American Journal of Gastroenterology, 93: 18031809.Google Scholar
Camilleri, M., Malagelada, J. R., Brown, M. L., Becker, G., & Zinsmeister, A. R. (1985). Relation between antral motility and gastric emptying of solids and liquids in humans. American Journal of Physiology, 249: G580G585.Google Scholar
Cannon, W. B. (1936). Digestion and Health. New York: Norton.Google Scholar
Cannon, W. B. & Washburn, A. L. (1912). An explanation of hunger. American Journal of Physiology, 29: 441454.Google Scholar
Carabotti, M., Scirocco, A., Maselli, M. A., & Severi, C. (2015). The gut–brain axis: interactions between enteric microbiota, central, and enteric nervous systems. Annals of Gastroenterology, 28: 203209.Google Scholar
Carlson, A. J. (1916). The Control of Hunger in Health and Disease. University of Chicago Press.Google Scholar
Chen, J. D. Z., Davenport, K., & McCallum, R. W. (1993a). Effects of fat preload on gastric myoelectrical activity in normal humans. Journal of Gastrointestinal Motility, 5: 281287.Google Scholar
Chen, J. D. Z., Lin, Z. Y., Parolisi, S., & McCallum, R. W. (1995). Inhibitory effects of cholecystokinin on postprandial gastric myoelectric activity. Digestive Diseases and Sciences, 40: 26142622.Google Scholar
Chen, J. D. Z. & McCallum, R. W. (1991). Electrogastrogram: measurement, analysis and prospective applications. Medical and Biological Engineering and Computing, 29: 339350.Google Scholar
Chen, J. D. Z., Pan, J., & Orr, W. C. (1996). Role of sham feeding in postprandial changes of gastric myoelectrical activity. Digestive Diseases and Sciences, 41: 17061712.Google Scholar
Chen, J. D. Z., Richards, R., & McCallum, R. W. (1993b). Frequency components of the electrogastrogram and their correlations with gastrointestinal motility. Medical and Biological Engineering and Computing, 31: 6067.Google Scholar
Chen, J. D. Z., Stewart, W. R., & McCallum, R. W. (1993c). Adaptive spectral analysis of episodic rhythmic variations in gastric myoelectric potentials. IEEE Transactions on Biomedical Engineering, 40: 128135.Google Scholar
Chen, J. D. Z., Vandewalle, J., Sansen, W., Vantrappen, G., & Janssens, J. (1990). Adaptive spectral analysis of cutaneous electrical signals using autoregressive moving average modeling. Medical and Biological Engineering and Computing, 28: 531536.Google Scholar
Chen, J. D. Z., Zou, X., Lin, X., Ouyang, S., & Liang, J. (1999). Detection of slow wave propagation from the cutaneous electrogastrogram. American Journal of Physiology, 277: G424G430.Google Scholar
Cheng, L. K., Du, P., & O’Grady, G. (2013). Mapping and modeling gastrointestinal bioelectricity: from engineering bench to bedside. Physiology, 28: 310317.Google Scholar
Chiloiro, M., Riezzo, G., Guerra, V., Reddy, S. N., & Girgio, I. (1994). The cutaneous electrogastrogram reflects postprandial gastric emptying in humans. In Chen, J. D. Z. & McCallum, R. W. (eds.), Electrogastrography: Principles and Applications (pp. 293306). New York: Raven Press.Google Scholar
Code, C. F. & Marlett, J. A. (1975). The interdigestive myo-electric complex of the stomach and small bowel of dogs. Journal of Physiology, 246: 289309.Google Scholar
Coelho-Aguiar, J. M., Bon-Frauches, A. C., Gomes, A. L. T., Verissimo, C. P., Aguiar, D. P., Matias, D., … & Moura-Neta, V. (2015). The enteric glia: identity and functions. Glia, 63: 921935.Google Scholar
Couturier, D., Roze, C., Paologgi, J., & Debray, C. (1972). Electrical activity of the normal human stomach: a comparative study of recordings obtained from serosal and mucosal sites. Digestive Diseases and Sciences, 17: 969976.Google Scholar
Davis, R. C., Garafolo, L., & Gault, F. P. (1957). An exploration of abdominal potentials. Journal of Comparative and Physiological Psychology, 50: 519523.Google Scholar
Davis, R. C., Garafolo, L., & Kveim, K. (1959). Conditions associated with gastrointestinal activity. Journal of Comparative and Physiological Psychology, 52: 466475.Google Scholar
Diamanti, A., Bracci, F., Gambarara, M., Ciofetta, G. C., Sabbi, T., Ponticelli, A., … & Castro, M. (2003). Gastric electric activity assessed by electrogastrography and gas emptying scintigraphy in adolescents with eating disorders. Journal of Pediatric Gastroenterology and Nutrition, 37: 3541.Google Scholar
Dickman, R., Zilper, T., Steinmetz, A., Pakanaev, L., Ron, Y., Bernstine, H., … & Shirin, H. (2013). Comparison of continuous breath test and gastric scintigraphy for the measurement of gastric emptying rate in healthy and dyspeptic individuals. European Journal of Gastroenterology and Hepatology, 25: 291295.Google Scholar
Dubois, M. & Mizrahi, M. (1994). Electrogastrography, gastric emptying, and gastric motility. In Chen, J. D. Z. & McCallum, R. W. (eds.). Electrogastrography: Principles and Applications (pp. 247256). New York: Raven Press.Google Scholar
Enck, P., Hefner, J, Herbert, B. M., Mazurak, N., Weimer, K., Muth, E. R., … & Martens, U. (2013). Sensitivity and specificity of hypnosis effects on gastric myoelectrical activity. PLoS One, 8: e83486.Google Scholar
Familoni, B. O., Bowes, K. L., Kingma, Y. J., & Cote, K. R. (1991). Can transcutaneous electrodes diagnose gastric electrical abnormalities? Gut, 32: 141146.Google Scholar
Friesen, C. A., Lin, Z., Schurman, J. V., Andre, L., & McCallum, R. W. (2007). Autonomic nervous system response to a solid meal and water loading in healthy children: its relation to gastric myoelectrical activity. Neurogastroenterology and Motility, 19: 376382.Google Scholar
Furness, J. B. & Costa, M. (1980). Types of nerves in the enteric nervous system. Neuroscience, 5: 120.Google Scholar
Gabbay, F. H. & Stern, R. M. (2012). A quiet voice: Roland Clark Davis and the emergence of psychophysiology. Psychophysiology, 49: 443453.Google Scholar
Geldof, H., van der Schee, E. J., & Grashuis, J. L. (1986a). Accuracy and reliability of electrogastrography (EGG). Gastroenterology, 90: 1425.Google Scholar
Geldof, H., van der Schee, E. J., van Blankenstein, M., & Grashuis, J. L. (1986b). Electrogastrographic study of gastric myoelectrical activity in patients with unexplained nausea and vomiting. Gut, 27: 799808.Google Scholar
Ghoos, Y. F., Maes, B. D., Geypens, B. J., Mys, G., Hiele, M. I., Rutgeerts, P. J., & Vantrappen, G. (1993). Measurement of gastric emptying rate of solids by means of a carbon-labeled octanoic acid breath test. Gastroenterology, 104: 16401647.Google Scholar
Gianaros, P. J., Quigley, K. S., Mordkoff, J. T., & Stern, R. M. (2001). Gastric myoelectrical and autonomic cardiac reactivity to laboratory stressors. Psychophysiology, 38: 642652.Google Scholar
Grover, M., Farrugia, G., Lurken, M. S., Bernard, C. E., Faussone-Pellegrini, M. S., Smyrk, T. C., … & Pasricha, P. J. (2011). Cellular changes in diabetic and idiopathic gastroparesis. Gastroenterology, 140: 15751585.Google Scholar
Hamilton, J. W., Bellahsene, B. E., Reichelderfer, M., Webster, J. H., & Bass, P. (1986). Human electrogastrograms: comparison of surface and mucosal recordings. Digestive Diseases and Sciences, 31: 3339.Google Scholar
Harrison, N. A., Gray, M. A., Gianaros, P. J., & Critchley, H. D. (2010). The embodiment of emotional feelings in the brain. Journal of Neuroscience, 30: 1287812884.Google Scholar
Hasler, W. L. (2014). The use of SmartPill for gastric monitoring. Expert Review of Gastroenterology and Hepatology, 8: 587600.Google Scholar
Herbert, B. M., Muth, E. R., Pollatos, O., & Herbert, C. (2012). Interoception across modalities: on the relationship between cardiac awareness and the sensitivity for gastric functions. PLoS One, 7: e36646.Google Scholar
Hinder, R. A. & Kelly, K. A. (1977). Human gastric pacesetter potentials: site of origin and response to gastric transection and proximal vagotomy. American Journal of Physiology, 133: 2933.Google Scholar
Hölzl, R., Loffler, K., & Muller, G. M. (1985). On conjoint gastrography or what the surface gastrograms show. In Stern, R. M. & Koch, K. L. (eds.), Electrogastrography: Methodology, Validation, and Applications (pp. 89115). New York: Praeger.Google Scholar
Huisinga, J. D. (2001). Physiology and pathophysiology of the interstitial cells of Cajal: from bench to bedside. II: Gastric motility: lessons from mutant mice on slow waves and innervation. American Journal of Physiology, 281: G1129G1134.Google Scholar
Jednak, M. A., Shadigian, E. M., Kim, M. S., Woods, M. L., Hooper, F. G., Owyang, , & Hasler, W. L. (1999). Protein meals reduce nausea and gastric slow wave dysrhythmic activity in first trimester pregnancy. American Journal of Physiology, 277: G855861.Google Scholar
Jokerst, M. D., Gatto, M., Fazio, R., Stern, R. M., & Koch, K. L. (1999). Slow deep breathing prevents the development of tachygastria and symptoms of motion sickness. Aviation, Space, and Environmental Medicine, 70: 11891192.Google Scholar
Jones, K. R. & Jones, G. E. (1985). Pre- and postprandial EGG variation. In Stern, R. M. & Koch, K. L. (eds.), Electrogastrography: Methodology, Validation and Applications (pp. 168181). New York: Praeger.Google Scholar
Keller, J., Andresen, V., Wolter, J., Layer, P., & Camilleri, M. (2009). Influence of clinical parameters on the results of 13C-ooctanoic acid breath tests: examination of different mathematical models in a large patient cohort. Neurogastroenterology and Motility, 21: 10391047.Google Scholar
Kelly, K. A., Code, C. F., & Elveback, L. R. (1969). Patterns of canine gastric electrical activity. American Journal of Physiology, 217: 461470.Google Scholar
Kim, J., Napadow, V., Kuo, B., & Barbieri, R. (2011). A combined HRV–fMRI approach to assess cortical control of cardiovagal modulation by motion sickness. In Proceedings of the IEEE Engineering in Medicine and Biological Society, 2011, (pp. 28252828). Piscataway, NJ: IEEE.Google Scholar
Kim, T. W., Beckett, E. A., Hanna, R., Koh, S. D., Ordög, T., Ward, S. M., & Sanders, K. M. (2002). Regulation of pacemaker frequency in the murine gastric antrum. Journal of Physiology, 538: 145157.Google Scholar
Kingma, Y. J. (1989). The electrogastrogram and its analysis. Critical Reviews in Biomedical Engineering, 17: 105124.Google Scholar
Koch, K. L. (2002). Electrogastrography. In Schuster, M. M., Crowell, M. D., & Koch, K. L. (eds.), Schuster Atlas of Gastrointestinal Motility, 2nd edn. (pp. 185202). Hamilton, ON: Decker.Google Scholar
Koch, K. L. (2014) Gastric dysrhythmias: a potential objective measure of nausea. Experimental Brain Research, 232: 25532561.Google Scholar
Koch, K. L., Hong, S. P., & Xu, L. (2000). Reproducibility of gastric myoelectrical activity and the water load test in patients with dysmotility-like dyspepsia symptoms and in control subjects. Journal of Clinical Gastroenterology, 12: 125129.Google Scholar
Koch, K. L. & Stern, R. M. (1985). The relationship between the cutaneously recorded electrogastrogram and antral contractions in man. In Stern, R. M. & Koch, K. L. (eds.), Electrogastrography: Methodology, Validation, and Applications (pp. 116131). New York: Praeger.Google Scholar
Koch, K. L, & Stern, R. M. (1993). Electrogastrography. In Kumar, D. & Wingate, D. (eds.), An Illustrated Guide to Gastrointestinal Motility (pp. 290307). London: Churchill Communications Europe.Google Scholar
Koch, K. L. & Stern, R. M. (2004). Handbook of Electrogastrography. Oxford University Press.Google Scholar
Koch, K. L., Stern, R. M., Bingaman, S., & Eggli, D. (1991). Satiety, stomach volume, and gastric myoelectrical activity during solid-phase gastric emptying: a study of healthy individuals. Journal of Gastrointestinal Motility, 5: 4147.Google Scholar
Koch, K. L., Stewart, W. R., & Stern, R. M. (1987). Effects of barium meals on gastric electromechanical activity in man: a fluorscopic-electrogastrophic study. Digestive Diseases and Sciences, 32: 12171222.Google Scholar
Koch, K. L., Tran, T. N., Stern, R. M., Bingaman, S., & Sperry, N. (1993). Gastric myoelectrical activity in premature and term infants. Neurogastroenterology & Motility, 5: 4147.Google Scholar
Kokubo, T., Matsui, S., & Ishiguro, M. (2013). Meta-analysis of oro-cecal transit time in fasting subjects. Pharmaceutical Research, 30: 402411.Google Scholar
Kundu, S. & Koch, K. L. (2011). Effect of balloon distention or botulinum toxin A injection of the pylorus on symptoms and body weight in patients with gastroparesis and normal 3 cycle per minute gastric electrical activity. American Journal of Gastroenterology, 106: S35.Google Scholar
Kwong, N. K., Brown, B. H., Whittaker, G. E., & Duthie, H. L. (1970). Electrical activity of the gastric antrum in man. British Journal of Surgery, 12: 913916.Google Scholar
Lacy, B. E., Koch, K. L., & Crowell, M. D. (2002). Manometry. In Schuster, M. M., Crowell, M. D., & Koch, K. L. (eds.), Schuster Atlas of Gastrointestinal Motility, 2nd edn. (pp. 135150). Hamilton, ON: Decker.Google Scholar
Lavigne, M. E., Wiley, Z. D., Meyer, J. H., Martin, P., & MacGregor, I. L. (1978). Gastric emptying rates of solid food in relation to body size. Gastroenterology, 74: 12581260.Google Scholar
Lee, K., Chey, W., Tai, H., & Yajima, H. (1978). Radioimmunoassay of motilin: validation and studies on the relationship between plasma motilin and interdigestive myoelectric activity in the duodenum of dog. Digestive Diseases and Sciences, 23: 789795.Google Scholar
Levine, M. E., DeRusso, D. M., Tehan, J. M., & Shafer, A. L. (2010). The beneficial effects of ginger for the prevention of motion-induced nausea and gastric dysrhythmia. Psychosomatic Medicine, 72: A109.Google Scholar
Levine, M. E., Holt, L., & Koch, K. L. (2007). Effects of soy, casein, and whey protein-predominant drinks on gastric myoelectrical activity and postprandial symptoms of fullness and satiety. Neurogastroenterology & Motility, 19: 526527.Google Scholar
Levine, M. E., Muth, E. R., Williamson, M. J., & Stern, R. M. (2004). Protein-predominant meals inhibit the development of gastric tachyarrhythmia, nausea and the symptoms of motion sickness. Alimentary Pharmacology & Therapeutics, 19: 583590.Google Scholar
Levine, M. E. & Puzino, K. M. (2013). Music helps but shadowing doesn’t: the effects of emotional and cognitive distraction on motion-induced nausea and gastric dysrhythmia. Psychosomatic Medicine, 75: A87.Google Scholar
Levine, M. E. & Stern, R. M. (2015). The use of facial cooling for the management of motion-induced nausea and gastric dysrhythmia. Psychosomatic Medicine, 77: A112.Google Scholar
Levine, M. E., Stern, R. M., & Koch, K. L. (2014). Enhanced perceptions of control and predictability reduce motion-induced nausea and gastric dysrhythmia. Experimental Brain Research, 232: 26752684.Google Scholar
Lin, H. C. & Hasler, W. L. (1995). Disorders of gastric emptying. In Yamada, T. (ed.), Textbook of Gastroenterology (pp. 13181346). Philadelphia, PA: J. P. Lippincott.Google Scholar
Lin, Z. & Chen, J. D. Z. (1994). Comparison of three running spectral analysis methods. In Chen, J. D. Z. & McCallum, R. W. (eds.), Electrogastrography: Principles and Applications (pp. 7598). New York: Raven Press.Google Scholar
Lunding, J. A., Nordstrom, L. M., Haukelid, A. O., Gilja, O. H., Berstad, A., & Hausken, T. (2008). Vagal activation by sham feeding improves gastric motility in functional dyspepsia. Neurogastroenterology and Motility, 20: 618624.Google Scholar
Marie, I., Gourcerol, G., Leroi, A., Menard, J., Levesque, H., & Ducrotte, P. (2012). Delayed gastric emptying determined using the 13C-octanoic acid breath test in patients with systemic sclerosis. Arthritis and Rheumatism, 64: 23462355.Google Scholar
Mayer, E. A., Tillisch, K., & Gupta, A. (2015). Gut/brain axis and the microbiota. Journal of Clinical Investigation, 125: 926938.Google Scholar
McCallum, R. W., Snape, W., Brody, F., Wo, J., Parkman, H. P., & Nowak, T. (2010). Gastric electrical stimulation with Enterra therapy improves symptoms from diabetic gastroparesis in a prospective study. Clinical Gastroenterology and Hepatology, 8: 947954.Google Scholar
McNearney, T., Lin, X., Shrestha, J., Lisse, J., & Chen, J. D. Z. (2002). Characterization of gastric myoelectrical rhythms in patients with systemic sclerosis using multichannel surface electrogastrography. Digestive Diseases & Sciences, 47: 690698.Google Scholar
Meissner, K. (2009). Effects of placebo interventions on gastric motility and general autonomic activity. Journal of Psychosomatic Research, 66: 391398.Google Scholar
Meyer, J. H. (1987). Motility of the stomach and gastroduodenal junction. In Johnson, L. R., Christensen, J., Jacobsen, E. D., & Schultz, S. G. (eds.), Physiology of the Gastrointestinal Tract (pp. 613630). New York: Raven Press.Google Scholar
Meyer, J. H., Gu, Y., Elashoff, J., Reedy, T., Dressman, J., & Amidon, G. (1986). Effect of viscosity and flow rate on gastric emptying of solids. American Journal of Physiology, 250: G161G164.Google Scholar
Meyer, J. H., MacGregor, I. L., Gueller, R., Martin, P., & Cavalieri, R. (1976). 99Tc-tagged chicken liver as a marker of solid food in the human stomach. American Journal of Digestive Diseases, 21: 296304.Google Scholar
Meyer, J. H., Ohashi, H., Jehn, D., & Thompson, J. B. (1981). Size of liver particles emptied from the human stomach. Gastroenterology, 80: 14891496.Google Scholar
Mintchev, M. P., Otto, S. J., & Bowes, K. L. (1997). Electrogastrography can recognize gastric electrical uncoupling in dogs. Gastroenterology, 112: 20062011.Google Scholar
Mirizzi, N. & Scafoglieri, V. (1983). Optimal direction of the electrogastrographic signal in man. Medical and Biological Engineering and Computing, 21: 385389.Google Scholar
Moraes, E. R., Toncon, L. E., Baffa, O., Oba-Kunyioshi, A. S., Wakai, R., & Leuthold, A. (2003). Adaptive, autoregressive spectral estimation for analysis of electrical signals of gastric origin. Physiological Measurement, 24: 91106.Google Scholar
Morgan, K. G., Schmalz, P. F., & Szurszewski, J. H. (1978). The inhibitory effects of vasoactive intestinal polypeptide on the mechanical and electrical activity of canine antral smooth muscle. Journal of Physiology, 282: 437450.Google Scholar
Muth, E. R., Koch, K. L, & Stern, R. M. (2000). Significance of autonomic nervous system activity in functional dyspepsia. Digestive Diseases and Sciences, 45: 854863.Google Scholar
Muth, E. R., Koch, K. L., Stern, R. M., & Thayer, J. F. (1999). Effect of autonomic nervous system manipulations on gastric myoelectrical activity and emotional responses in healthy human subjects. Psychosomatic Medicine, 61: 297303.Google Scholar
Muth, E. R., Stern, R. M., & Koch, K. L. (1996). Effects of vection-induced motion sickness on gastric myoelectric activity and oral-cecal transit time. Digestive Diseases and Sciences, 41: 330334.Google Scholar
Napadow, V., Sheehan, J., Kim, J., Dassatti, A., Thurler, A. H., Surjanhata, B., … & Kuo, B. (2013a). Brain white matter microstructure is associated with susceptibility to motion-induced nausea. Neurogastroenterology and Motility, 25: 448450.Google Scholar
Napadow, V., Sheehan, J. D., Kim, J., LaCount, L. T., Park, K., Kaptchuk, T. J., … & Kuo, B. (2013b). The brain circuitry underlying the temporal evolution of nausea in humans. Cerebral Cortex, 23: 806813.Google Scholar
Nelsen, T. S. & Kohatsu, S. (1968). Clinical electrogastrography and its relationship to gastric surgery. American Journal of Surgery, 116: 215222.Google Scholar
Neunlist, M., Rolli-Derkinderen, M., Latorre, R., Van Landeghem, L., Coron, E., Derkinderen, P., & DeGiorgio, R. (2014). Enteric glial cells: recent developments and future directions. Gastroenterology, 147: 12301237.Google Scholar
Nobrega, A. C. M., Ferreira, B. R. S., Oliveira, G. J., Sales, K. M. O., Santos, A. A., Nobre E Souza, M. A., … & Souza, M. H. L. P. (2012). Dyspeptic symptoms and delayed gastric emptying of solids in patients with inactive Crohn’s disease. BMC Gastroenterology, 12: 175.Google Scholar
Nonaka, T., Inamori, M., Endo, H., Matsuura, M., Uchiyama, S., Yamada, E., … & Maeda, S. (2014). Correlation between gastric transit time measured by video capsule endoscopy and gastric emptying determined by the continuous real-time 13C breath test (Breath ID system). Hepatogastroenterology, 61: 21592162.Google Scholar
Ogawa, A., Mizuta, I., Fukunaga, T., Takeuchi, N., Honaga, E., Sugita, Y., … & Takeda, M. (2004). Electrogastrography abnormality in eating disorders. Psychiatry and Clinical Neurosciences, 58: 300310.Google Scholar
O’Grady, G., Angeli, T., Du, P., Lahr, C., Lammers, W. J., Windsor, J. A., … & Cheng, L. K. (2012). Abnormal initiation and conduction of slow wave activity in gastroparesis defined by a high resolution electrical mapping. Gastroenterology, 143: 589598.Google Scholar
Peyrot des Gachons, C., Beauchamp, G. K., Stern, R. M., Koch, K. L., & Breslin, P. A. (2011). Bitter taste induces nausea. Current Biology, 21: R247R248.Google Scholar
Qin, S., Miao, L., Xi, N., Wang, Y., & Yang, C. (2010). A real-time weighted-eigenvector MUSIC method for time-frequency analysis of electrogastrogram slow wave. In Proceedings of the IEEE Engineering in Medicine and Biological Society, 2010 (pp. 867870). Piscataway, NJ: IEEE.Google Scholar
Read, N. W., An-Janabi, M. N., Bates, T. E., Holgate, A. M., Cann, P. A., Kinsman, R. I., … & Brown, C. (1985). Interpretation of the breath hydrogen profile obtained after ingesting a solid meal containing unabsorbable carbohydrate. Gut, 26: 834842.Google Scholar
Riezzo, G., Castellana, R. M., De Bellis, T., Laforgia, F., Indrio, F., & Chilorio, M. (2003). Gastric electrical activity in normal neonates during the first year of life: effect of feeding with breast milk and formula. Journal of Gastroenterology, 38: 836843.Google Scholar
Riezzo, G., Porcelli, P., Guerra, V., & Giorgio, I. (1996). Effects of different psychophysiological stressors on the cutaneous electrogastrogram in healthy subjects. Archives of Physiology and Biochemistry, 104: 282286.Google Scholar
Roman, C. & Gonella, J. (1987). Extrinsic control of digestive tract motility. In Johnson, L. R., Christensen, J., Jacobsen, E. D., & Schultz, S. G. (eds.), Physiology of the Gastrointestinal Tract (pp. 507553). New York: Raven Press.Google Scholar
Sarna, S. K. (2002). Myoelectrical and contractile activities of the gastrointestinal tract. In Schuster, M. M., Crowell, M. D., & Koch, K. L. (eds.), Schuster Atlas of Gastrointestinal Motility, 2nd edn. (pp. 118). Hamilton, ON: Decker.Google Scholar
Schlegel, J. F. & Code, C. F. (1975). The gastric peristalsis of the interdigestive housekeeper. In Vantrappen, G. (ed.), Proceedings from the Fifth International Symposium on Gastrointestinal Motility (p. 321). Herentals, Belgium: Typoff Press.Google Scholar
Sciarretta, G., Furno, A., Mazzoni, M., Garagnani, B., & Malagut, P. (1994). Lactulose hydrogen breath test in orocecal transit assessment: critical evaluation by means of scintigraphic method. Digestive Diseases and Sciences, 39: 15051510.Google Scholar
Sclocco, R., Kim, J., Garcia, R. G., Sheehan, J. D., Beissner, F., Bianchi, A. M., … & Napadow, V. (2016). Brain circuitry supporting multi-organ autonomic outflow in response to nausea. Cerebral Cortex, 26: 485497.Google Scholar
Scott, B. K., Koch, K. L., & Westcott, C. J. (2014). Pyloroplasty for patients with medically-refractory functional obstructive gastroparesis. Gastroenterology, 146: S615.Google Scholar
Smallwood, R. H. (1978). Analysis of gastric electrical signals from surface electrodes using phase-lock techniques. Part 2: System performance with gastric signals. Medical and Biological Engineering and Computing, 16: 513518.Google Scholar
Smallwood, R. H. & Brown, B. H. (1983). Non-invasive assessment of gastric activity. In Rolfe, P. (ed.), Non-Invasive Physiological Measurements, Volume 2. London: Academic Press.Google Scholar
Smout, A. J. P. M. (1980). Myoelectric activity of the stomach: gastroelectromyography and electrogastrography. Unpublished thesis, Erasmus University, Rotterdam.Google Scholar
Smout, A. J. P. M., Jebbink, H. J. A., & Samsom, M. (1994). Acquisition and analysis of electrogastrographic data: the Dutch experience. In Chen, J. D. Z. & McCallum, R. W. (eds.), Electrogastrography: Principles and Applications (pp. 330). New York: Raven Press.Google Scholar
Smout, A. J. P. M., van der Schee, E. J., & Grashuis, J. L. (1980a). Postprandial and interdigestive gastric electrical activity in the dog recorded by means of cutaneous electrodes. In Christensen, J. (ed.), Gastrointestinal Motility (pp. 187194). New York: Raven Press.Google Scholar
Smout, A. J. P. M., van der Schee, E. J., & Grashuis, J. L. (1980b). What is measured in electrogastrography? Digestive Diseases and Sciences, 25: 179187.Google Scholar
Stemper, T. J. & Cooke, A. R. (1975). Gastric emptying and its relationship to antral contractile activity. Gastroenterology, 69: 649653.Google Scholar
Stern, R. M., Crawford, H. E., Stewart, W. R., Vasey, M. W., & Koch, K. L. (1989). Sham feeding: cephalic-vagal influences on gastric myoelectric activity. Digestive Diseases and Sciences, 34: 521527.Google Scholar
Stern, R. M., Jokerst, M. D., Levine, M. E., & Koch, K. L. (2001). The stomach’s response to unappetizing food: cephalic-vagal effects on gastric myoelectric activity. Neurogastroenterology and Motility, 13: 151154.Google Scholar
Stern, R. M. & Koch, K. L. (1994). Using the electrogastrogram to study motion sickness. In Chen, J. D. Z. & McCallum, R. W. (eds.), Electrogastrography: Principles and Applications. (pp. 199218). New York: Raven Press.Google Scholar
Stern, R. M. & Koch, K. L. (1996). Motion sickness and differential susceptibility. Current Directions in Psychological Science, 5: 115120.Google Scholar
Stern, R. M., Koch, K. L., & Andrews, P. L. R. (2011). Nausea: Mechanisms and Management. Oxford University Press.Google Scholar
Stern, R. M., Koch, K. L., Leibowitz, H. W., Lindblad, I., Shupert, C., & Stewart, W. R. (1985). Tachygastria and motion sickness. Aviation Space and Environmental Medicine, 56: 10741077.Google Scholar
Stern, R. M., Koch, K. L., & Muth, E. R. (2000). Gastrointestinal system. In Cacioppo, J. T., Tassinary, L. G., & Berntson, G. G. (eds.), Handbook of Psychophysiology, 2nd edn. (pp. 294314). Cambridge University Press.Google Scholar
Stern, R. M., Koch, K. L., Stewart, W. R., & Lindblad, I. M. (1987). Spectral analysis of tachygastria recorded during motion sickness. Gastroenterology, 92: 9297.Google Scholar
Stern, R. M. & Stacher, G. (1982). Recording the electrogastrogram from parts of the body surface distant from the stomach. Psychophysiology, 19: 350.Google Scholar
Stern, R. M., Vasey, M. W., Hu, S., & Koch, K. L. (1991). Effects of cold stress on gastric myoelectric activity. Journal of Gastrointestinal Motility, 3: 225228.Google Scholar
Stern, R. M., Vitellaro, K., Thomas, M., Higgins, S. C., & Koch, K. L. (2004). Electrogastrographic biofeedback: a technique for enhancing normal gastric activity. Neurogastroenterology and Motility, 16: 753757.Google Scholar
Stoddard, C. J., Smallwood, R. H., & Duthie, H. L. (1981). Electrical arrhythmias in the human stomach. Gut, 22: 705712.Google Scholar
Thompson, D. G., Richelson, E., & Malagelada, J. R. (1982). Perturbation of gastric emptying and duodenal motility through the central nervous system. Gastroenterology, 83: 12001206.Google Scholar
Thunberg, L. (1989). Interstitial cells of Cajal. In Wood, J. D. (ed..), Handbook of Physiology: The Gastrointestinal System (pp. 349386). Bethesda, MD: American Physiological Society.Google Scholar
Uijtdehaage, S. H. J., Stern, R. M., & Koch, K. L. (1992) Effects of eating on vection-induced motion sickness, cardiac vagal tone and gastric myoelectric activity. Psychophysiology, 29: 193201.Google Scholar
Van der Schee, E. J., Smout, A. J. P. M., & Grashuis, J. L. (1982). Applications of running spectrum analysis to electrogastrographic signals recorded from dog and man. In Wienbeck, M. (ed.), Motility of the Digestive Tract (pp. 241250). New York: Raven Press.Google Scholar
Vantrappen, G., Hostein, J., Janssens, J., Vanderweerd, M., & De Wever, I. (1983). Do slow waves induce mechanical activity?Gastroenterology, 84: 1341.Google Scholar
Vantrappen, G., Janssens, J., Peeters, T. L., Bloom, S., R., Christofides, N. D., & Hellemans, J. (1979). Motility and the interdigestive migrating motor complex in man. Digestive Diseases and Sciences, 24: 497500.Google Scholar
Verbeke, K. (2009). Will the 13C-octanoic acid breath test ever replace scintigraphy as the gold standard to assess gastric emptying? Neurogastroenterology and Motility, 21: 10131016.Google Scholar
ver Hagen, M. A. M. T., Luijk, H. D., Samsom, M., & Smout, A. J. P. M. (1998). Effect of meal temperature on the frequency of gastric myoelectrical activity. Neurogastroenterolog and Motility, 10: 175181.Google Scholar
Vianna, E. P. M., Naqvi, N., Bechara, A., & Tranel, D. (2009). Does vivid emotional imagery depend on body signals? International Journal of Psychophysiology, 72: 4650.Google Scholar
Vianna, E. P. M. & Tranel, D. (2006). Gastric myoelectrical activity as an index of emotional arousal. International Journal of Psychophysiology, 61: 7076.Google Scholar
Vianna, E. P. M., Weinstock, J., Elliott, D., Summers, R., & Tranel, D. (2006). Increased feelings with increased body signals. Social, Cognitive, and Affective Neuroscience, 1: 3748.Google Scholar
Wolf, S. (1943). Relation of gastric function to nausea in man. Journal of Clinical Investigations, 22: 877882.Google Scholar
Wolf, S. & Wolff, H. G. (1943). Human Gastric Function. Oxford University Press.Google Scholar
Wood, J. D. (2002). Neural and humoral regulation of gastrointestinal motility. In Schuster, M. M., Crowell, M. D., & Koch, K. L. (eds.), Atlas of Gastrointestinal Motility, 2nd edn. (pp. 1942). Hamilton, ON: Decker.Google Scholar
Xu, L., Koch, K. L., Gianaros, P. J., Schreibman, I. R., Ku, M., & Rolls, B. J. (2002). The effects of soup, casserole, and water ingestion on gastric myoelectrical activity and perception of hunger and fullness. Gastroenterology, 122: A326.Google Scholar
Xu, X., Chen, D. D., Yin, J., & Chen, J. D. Z. (2014). Altered postprandial responses in gastric myoelectrical activity and cardiac autonomic functions in healthy obese subjects. Obesity Surgery, 24: 554560.Google Scholar
Yin, J. & Chen, J. D. Z. (2013). Electrogastrography: methodology, validation, and applications. Journal of Neurogastroenterology and Motility, 19: 517.Google Scholar
You, C. H. & Chey, W. Y. (1984). Study of electromechanical activity of the stomach in humans and in dogs with particular attention to tachygastria. Gastroenterology, 86: 14601468.Google Scholar

References

American Psychiatric Association [APA] (1980). Diagnostic and Statistical Manual of Mental Disorders III (DSM-III). Arlington, VA: APA.Google Scholar
Andersson, K. E. (2011). Mechanisms of penile erection and basis for pharmacological treatment of erectile dysfunction. Pharmacological Review, 63: 811859.Google Scholar
Arnow, B. A., Desmond, J. E., Banner, L. L., Glover, G. H., Solomon, A., Polan, M. L., … & Atlas, S. W. (2002). Brain activation and sexual arousal in healthy, heterosexual males. Brain, 125: 10141023.Google Scholar
Arnow, B. A., Millheiser, L., Garrett, A., Lake Polan, M., Glover, G. H., Hill, K. R., … & Desmond, J. E. (2009). Women with hypoactive sexual desire disorder compared to normal females: a functional magnetic resonance imaging study. Neuroscience, 158: 484502.Google Scholar
Bancroft, J. (1995). Are the effects of androgens on male sexuality noradrenergically mediated? Some consideration of the human. Neuroscience & Biobehavioral Reviews, 2: 16.Google Scholar
Bancroft, J. & Janssen, E. (2000). The dual control model of male sexual response: a theoretical approach to centrally mediated erectile dysfunction. Neuroscience & Biobehavioral Reviews, 24: 571579.Google Scholar
Bancroft, J., Jones, H. G., & Pullan, B. P. (1966). A simple transducer for measuring penile erection with comments on its use in the treatment of sexual disorder. Behavior Research and Therapy, 4: 239241.Google Scholar
Barlow, D. H. (1986). Causes of sexual dysfunction: the role of anxiety and cognitive interference. Journal of Consulting and Clinical Psychology, 54: 140157.Google Scholar
Barlow, D. H., Becker, R., Leitenberg, H., & Agras, W. (1970). A mechanical strain gauge for recording penile circumference change. Journal of Applied Behavior Analysis, 6: 355367.Google Scholar
Basson, R. (2002). A model of women’s sexual arousal. Journal of Sex and Marital Therapy, 28: 110.Google Scholar
Berns, G. S., McClure, S. M., Pagnoni, G., & Montague, P. R. (2001). Predictability modulates human brain response to reward. Journal of Neuroscience, 21: 27932798.Google Scholar
Bohlen, J. G. & Held, J. P. (1979). An anal probe for monitoring vascular and muscular events during sexual responsePsychophysiology, 16: 318323.Google Scholar
Borg, C., Georgiadis, J. R., Renken, R. J., Spoelstra, S. K., Weijmar Schultz, W., & de Jong, P. J. (2014). Brain processing of visual stimuli representing sexual penetration versus core and animal-reminder disgust in women with lifelong vaginismus. PLoS One, 9: e84882.Google Scholar
Bossio, J. A., Suschinsky, K. D., Puts, D. A., & Chivers, M. L. (2014). Does menstrual cycle phase influence the gender specificity of heterosexual women’s genital and subjective sexual arousal? Archives of Sexual Behavior, 43: 941952.Google Scholar
Both, S., van Lunsen, R., Weijenborg, P., & Laan, E. (2012). A new device for simultaneous measurement of pelvic floor muscle activity and vaginal blood flow: a test in a nonclinical sample. Journal of Sexual Medicine, 9: 28882902.Google Scholar
Bradley, W. E., Timm, G. W., Gallagher, J. M., & Johnson, B. K. (1985). New method for continuous measurement of nocturnal penile tumescence and rigidity. Urology, 26: 49.Google Scholar
Brom, M., Laan, E., Everaerd, W., Spinhoven, P., Cousijn, J., & Both, S. (2015). The influence of emotion down-regulation on the expectation of sexual reward. Behavior Therapy, 46: 379394.Google Scholar
Brotto, L. A. & Yule, M. A. (2011). Physiological and subjective sexual arousal in self-identified asexual women. Archives of Sexual Behavior, 40: 699712.Google Scholar
Burnett, A. L. (1997). Nitric oxide in penis: physiology and pathology. Journal of Urology, 157: 320324.Google Scholar
Burri, A., Heinrichs, M., Schedlowski, M., & Kruger, T. H. (2008). The acute effects of intranasal oxytocin administration on endocrine and sexual function in males. Psychoneuroendocrinology, 33: 591600.Google Scholar
Buvat, J. & Bou Jaoude, G. (2006). Significance of hypogonadism in erectile dysfunction. World Journal of Urology, 24: 657667.Google Scholar
Buvat, J., Maggi, M., Guay, A., & Torres, L. O. (2013). Testosterone deficiency in men: systematic review and standard operating procedures for diagnosis and treatment. Journal of Sexual Medicine, 10: 245284.Google Scholar
Cacioppo, S., Bianchi-Demicheli, F., Frum, C., Pfaus, J. G., & Lewis, J. W. (2012). The common neural bases between sexual desire and love: a multilevel kernel density fMRI analysis. Journal of Sexual Medicine, 9: 10481054.Google Scholar
Carmichael, M. S., Warburton, V. L., Dixen, J., & Davidson, J. M. (1994). Relationships among cardiovascular, muscular, and oxytocin responses during human sexual activity. Archives of Sexual Behavior, 23: 5979.Google Scholar
Carretié, L., Hinojosa, J. A., & Mercado, F. (2003). Cerebral patterns of attentional habituation to emotional visual stimuli. Psychophysiology, 40: 381388.Google Scholar
Carter, C. S. (1992). Oxytocin and sexual behavior. Neuroscience & Biobehavioral Reviews, 16: 131144.Google Scholar
Chivers, M. L., Seto, M. C., Lalumiere, M. L., Laan, E., & Grimbos, T. (2010). Agreement of self-reported and genital measures of sexual arousal in men and women: a meta-analysis. Archives of Sexual Behavior, 39: 556.Google Scholar
Clifton, J., Seehuus, M., & Rellini, A. H. (2015). Testing cognitive predictors of individual differences in the sexual psychophysiological responses of sexually functional women. Psychophysiology, 52: 957968.Google Scholar
Curtin, J. J., Lozano, D., & Allen, J. B. (2007). The psychophysiology laboratory. In Coan, J. A. & Allen, J. J. B. (eds.), Handbook of Emotion Elicitation and Assessment (pp. 398425). Oxford University Press.Google Scholar
Dawson, S. J., Sawatsky, M. L., & Lalumiere, M. L. (2015). Assessment of introital lubrication. Archives of Sexual Behavior, 44: 15271535.Google Scholar
Delizonna, L. L., Wincze, J. P., Litz, B. T., Brown, T. A., & Barlow, D. H. (2001). A comparison of subjective and physiological measures of mechanically produced and erotically produced erections. Journal of Sex & Marital Therapy, 27: 2131.Google Scholar
Dickinson, R. L. (1933). Human Sex Anatomy. Baltimore, MD: Williams & Wilkins.Google Scholar
Di Marino, V. & Lepidi, H. (2014). Microscopic study of the bulbo-clitoral organ. In Di Marino, V. & Lepidi, H., Anatomic Study of the Clitoris and the Bulbo-Clitoral Organ (pp. 5766). Dordrecht: Springer.Google Scholar
Earls, C. M., Marshall, W. L., Marshall, P. G., Morales, A., & Surridge, D. H., 1983. Penile elongation: a method for the screening of impotence. Journal of Urology, 139: 9092.Google Scholar
Everaerd, W., 1988. Commentary on sex research: sex as an emotion. Journal of Psychology and Human Sexuality, 2: 315.Google Scholar
Fernández-Cuevas, I., Bouzas Marins, J. C., Arnáiz Lastras, J., Gómez Carmona, P. M., Piñonosa Cano, S., García-Concepción, M. Á., & Sillero-Quintana, M. (2015). Classification of factors influencing the use of infrared thermography in humans: a review. Infrared Physics & Technology, 71: 2855.Google Scholar
Ferretti, A., Caulo, M., Del Gratta, C., Di Matteo, R., Merla, A., Montorsi, F., … & Romani, G. L. (2005). Dynamics of male sexual arousal: distinct components of brain activation revealed by fMRI. NeuroImage, 26: 10861096.Google Scholar
Fisher, C., Gross, J., & Zuch, J. (1965). Cycle of penile erection synchronous with dreaming (REM) sleep. Archives of General Psychiatry, 12: 2745.Google Scholar
Fisher, S. (1973). The Female Orgasm. New York: Basic Books.Google Scholar
Fisher, S. & Osofsky, H. (1967). Sexual responsiveness in women: psychological correlates. Archives of General Psychiatry, 17: 214226.Google Scholar
Freund, K. (1963). A laboratory method for diagnosing predominance of homo- or hetero-erotic interest in the male. Behaviour Research and Therapy, 1: 8593.Google Scholar
Freund, K., Langevin, R., & Barlow, D. (1974). Comparison of two penile measures of erotic arousal. Behaviour Research and Therapy, 12: 355359.Google Scholar
Fugl-Meyer, A. R., Sjogren, K., & Johansson, K. (1984). A vaginal temperature registration system. Archives of Sexual Behavior, 13: 247260.Google Scholar
Geer, J. H. (2005). Development of the vaginal photoplethysmograph. International Journal of Impotence Research, 17: 285287.Google Scholar
Geer, J. H. & Manguno, G. M. (1997). Gender differences in cognitive processes in sexuality. Annual Review of Sex Research, 9: 90124.Google Scholar
Georgiadis, J. R., Farrell, M. J., Boessen, R., Denton, D. A., Gavrilescu, M., Kortekaas, R., … & Egan, G. F. (2010). Dynamic subcortical blood flow during male sexual activity with ecological validity: a perfusion fMRI study. NeuroImage, 50: 208216.Google Scholar
Georgiadis, J. R. & Kringelbach, M. L. (2012). The human sexual response cycle: brain imaging evidence linking sex to other pleasures. Progress in Neurobiology, 98: 4981.Google Scholar
Georgiadis, J. R., Kringelbach, M. L., & Pfaus, J. G. (2012). Sex for fun: a synthesis of human and animal neurobiology. National Review of Urology, 9: 486498.Google Scholar
Gillath, O., Mikulincer, M., Birnbaum, G., & Shaver, P. R. (2007). Does subliminal exposure to sexual stimuli have the same effects on men and women? Journal of Sex Research, 44: 111.Google Scholar
Graber, B., Rohrbaugh, J., Newlin, D., Varner, J., & Ellingson, R. (1985). EEG during masturbation and ejaculation. Archives of Sexual Behavior, 14: 491503.Google Scholar
Graham, C. A., Sanders, S. A., Milhausen, R., & McBride, K. (2004). Turning on and turning off: a focus group study of the factors that affect women’s sexual arousal. Archives of Sexual Behavior, 33: 527538.Google Scholar
Gregory, R., Cheng, H., Rupp, H. A., Sengelaub, D. R., & Heiman, J. R. (2015). Oxytocin increases VTA activation to infant and sexual stimuli in nulliparous and postpartum women. Hormones and Behavior, 69: 8288.Google Scholar
Guillory, S. A. & Bujarski, K. A. (2014). Exploring emotions using invasive methods: review of 60 years of human intracranial electrophysiology. Social Cognitive and Affective Neuroscience, 9: 18801889.Google Scholar
Halpern, C. T, Udry, J. R, & Suchindran, C. (1998). Monthly measures of salivary testosterone predict sexual activity in adolescent males. Archives of Sexual Behavior, 27: 445465.Google Scholar
Hanson, R. K. & Bussiere, M. T. (1998). Predicting relapse: a meta-analysis of sexual offender recidivism studies. Journal of Consulting and Clinical Psychology, 66: 348362.Google Scholar
Heiman, J. R. (1977). A psychophysiological exploration of sexual arousal patterns in females and males. Psychophysiology, 14: 266274.Google Scholar
Helpman, L., Greenstein, A., Hartoov, J., & Abramov, L. (2009). Genito-sensory analysis in women with arousal and orgasmic dysfunction. Journal of Sexual Medicine, 6: 10391044.Google Scholar
Henson, D. E. & Rubin, H. B. (1978). A comparison of two objective measures of sexual arousal of women. Behaviour Research and Therapy, 16: 143151.Google Scholar
Hoon, P. W., Wincze, J. P., & Hoon, E. F. (1976). Physiological assessment of sexual arousal in women. Psychophysiology, 13: 196204.Google Scholar
Howett, M. K., Neely, E. B., Christensen, N. D., Wigdahl, B., Krebs, F. C., Malamud, D., … & Kreider, J. W. (1999). A broad-spectrum microbicide with virucidal activity against sexually transmitted viruses. Antimicrobial Agents and Chemotherapy, 43: 314321.Google Scholar
Huberman, J. S., Suschinsky, K. D., Lalumiere, M. L., & Chivers, M. L. (2013). Relationship between impression management and three measures of women’s self-reported sexual arousal. Canadian Journal of Behavioural Science / Revue Canadienne des Sciences du Comportement, 45: 259273.Google Scholar
Janssen, E. (2002). Psychophysiological measures of sexual response. In Wiederman, M. W. & Whitley, B. E. (eds.), Handbook for Conducting Research on Human Sexuality (pp. 139171). Mahwah, NJ: Lawrence Erlbaum Associates.Google Scholar
Janssen, E. (2011). Sexual arousal in men: a review and conceptual analysis. Hormones and Behavior, 59: 708716.Google Scholar
Janssen, E. & Bancroft, J. (2007). The dual-control model: the role of sexual inhibition and excitation in sexual arousal and behavior. In Janssen, E. (ed.), The Psychophysiology of Sex (pp. 197222). Bloomington, IN: Indiana University Press.Google Scholar
Janssen, E. & Everaerd, W. (1993). Determinants of male sexual arousal. Annual Review of Sex Research, 4: 211245.Google Scholar
Janssen, E., Everaerd, W., Spiering, M., & Janssen, J. (2000). Automatic processes and the appraisal of sexual stimuli: toward an information processing model of sexual arousal. Journal of Sex Research, 37: 823.Google Scholar
Janssen, E., McBride, K., Yarber, W., Hill, B. J., & Butler, S. (2008). Factors that influence sexual arousal in men: a focus group study. Archives of Sexual Behavior, 37: 252265.Google Scholar
Janssen, E., Vissenberg, M., Visser, S., & Everaerd, W. (1997). An in vivo comparison of two circumferential penile strain gauges: introducing a new calibration method. Psychophysiology, 34: 717720.Google Scholar
Janssen, E., Vorst, H., Finn, P., & Bancroft, J. (2002). The Sexual Inhibition (SIS) and Sexual Excitation (SES) Scales: II. Predicting psychophysiological response pattens. Journal of Sex Research, 39: 127132.Google Scholar
Kaplan, H. S. (1977). Hypoactive sexual desire. Journal of Sex and Marital Therapy, 3: 39.Google Scholar
Kaplan, H. S. (1979). Disorders of Sexual Desire. New York: Brunner/Mazel.Google Scholar
Karacan, I., Rosenbloom, A., & Williams, R. L. (1970). The clitoral erection cycle during sleep. Psychophysiology, 7: 338.Google Scholar
Karacan, I., Salis, P. J., Ware, J. C., Dervent, B., Williams, R. L., Scott, F. B., … & Beutler, L. E. (1978). Nocturnal penile tumescence and diagnosis in diabetic impotence. American Journal of Psychiatry, 135: 191197.Google Scholar
Khalifé, S., Binik, Y. M., Cohen, D. R., & Amsel, R. (2000). Evaluation of clitoral blood flow by color Doppler ultrasonography. Journal of Sex & Marital Therapy, 26: 187189.Google Scholar
Kinsey, A. C., Pomeroy, W. B., Martin, C. E., & Gebhardt, P. H. (1953). Sexual Behavior in the Human Female. Philadelphia: W. B. Saunders.Google Scholar
Knack, N. M., Murphy, L., Ranger, R., Meston, C., & Fedoroff, J. P. (2015). Assessment of female sexual arousal in forensic populations. Current Psychiatry Reports, 17: 557.Google Scholar
Komisaruk, B. R., Wise, N., Frangos, E., Liu, W. C., Allen, K., & Brody, S. (2011). Women’s clitoris, vagina, and cervix mapped on the sensory cortex: fMRI evidence. Journal of Sexual Medicine, 8: 28222830.Google Scholar
Krug, R., Plihal, W., Fehm, H. L., & Born, J. (2000). Selective influence of the menstrual cycle on perception of stimuli with reproductive significance: an event-related potential study. Psychophysiology, 37: 111122.Google Scholar
Kuban, M., Barbaree, H. E., & Blanchard, R. (1999). A comparison of volume and circumference phallometry: response magnitude and method agreement. Archives of Sexual Behavior, 28: 345359.Google Scholar
Kukkonen, T. M., Binik, Y. M., Amsel, R., & Carrier, S. (2007). Thermography as a physiological measure of sexual arousal in both men and women. Journal of Sexual Medicine, 4: 93105.Google Scholar
Kukkonen, T. M., Binik, Y. M., Amsel, R., & Carrier, S. (2010). An evaluation of the validity of thermography as a physiological measure of sexual arousal in a non-university adult sample. Archives of Sexual Behavior, 39(4), 861873.Google Scholar
Laan, E. & Everaerd, W. (1995). Determinants of female sexual arousal: psychophysiological theory and data. Annual Review of Sex Research, 6: 3276.Google Scholar
Laan, E., Everaerd, W., & Evers, A. (1995). Assessment of female sexual arousal: response specificity and construct validity. Psychophysiology, 32: 476485.Google Scholar
Lang, P. (1984). Cognition in emotion: concept and action. In Izard, C. E., Kagan, J., & Zajonc, R. B. (eds.), Emotions, Cognitions and Behavior. Cambridge University Press, 192225.Google Scholar
Lankveld, J. & Hout, M. A. (2004). Increasing neutral distraction inhibits genital but not subjective sexual arousal of sexually functional and dysfunctional men. Archives of Sexual Behavior, 33: 549558.Google Scholar
Larsen, J. T. & McGraw, A. P. (2014). The case for mixed emotions. Social and Personality Psychology Compass, 8: 263274.Google Scholar
Lawrence, A. A., Latty, E. M., Chivers, M. L., & Bailey, J. M. (2005). Measurement of sexual arousal in postoperative male-to-female transsexuals using vaginal photoplethysmography. Archives of Sexual Behavior, 34: 135145.Google Scholar
Levin, R. J. (1992). The mechanisms of human female sexual arousal. Annual Review of Sex Research, 3: 148.Google Scholar
Levin, R. J. (2003). A journey through two lumens! International Journal of Impotence Research, 15: 29.Google Scholar
Levin, R. J. (2015). Recreation and procreation: a critical view of sex in the human female. Clinical Anatomy, 28: 339354.Google Scholar
Levin, R. J. & Wagner, G. (1978). Haemodynamic changes of the human vagina during sexual arousal assessed by a heated oxygen electrode. Journal of Physiology, 275: 2324.Google Scholar
Levine, L. A. & Carroll, R. A. (1994). Nocturnal penile tumescence and rigidity in men without complaints of erectile dysfunction using a new quantitative analysis software. Journal of Urology, 152: 11031107.Google Scholar
Lindquist, K. A., Satpute, A. B., Wager, T. D., Weber, J., & Barrett, L. F. (2016). The brain basis of positive and negative affect: evidence from a meta-analysis of the human neuroimaging literature. Cerebral Cortex, 26: 19101922.Google Scholar
Lowenstein, L., Davis, C., Jesse, K., Durazo-Arvizu, R., & Kenton, K. (2009). Comparison between sensory testing modalities for the evaluation of afferent nerve functioning in the genital area. International Urogynecology Journal, 20: 8387.Google Scholar
Lübke, K. T., Hoenen, M., & Pause, B. M. (2012). Differential processing of social chemosignals obtained from potential partners in regards to gender and sexual orientation. Behavioural Brain Research, 228: 375387.Google Scholar
Marshall, W. L. (2014). Phallometric assessments of sexual interests: an update. Current Psychiatry Reports, 16: 428.Google Scholar
Masters, W. H. & Johnson, V. E. (1966). Human Sexual Response. New York: Little, Brown and Company.Google Scholar
McConaghy, N. (1974). Measurements of change in penile dimensions. Archives of Sexual Behavior, 3: 381388.Google Scholar
McKenna, K. E. (2000). Some proposals regarding the organization of the central nervous system control of penile erectionNeuroscience & Biobehavioral Reviews, 24: 535540.Google Scholar
Mendelsohn, M. (1896). Ist das Radfahren als eine gesundheidsgemässe uebung anzusehen und aus ärtzlichen gesichtspunkten zu empfehlen? Deutsche Medicinische Wochenschrift, 22: 383384.Google Scholar
Meston, C. M. & Gorzalka, B. B. (1995). The effects of sympathetic activation on physiological and subjective sexual arousal in women. Behaviour Research and Therapy, 3: 651664.Google Scholar
Meston, C. M., Rellini, A. H., & McCall, K. (2010). The sensitivity of continuous laboratory measures of physiological and subjective sexual arousal for diagnosing women with sexual arousal disorder. Journal of Sexual Medicine, 7: 938950.Google Scholar
Meuwissen, I. & Over, R. (1993). Female sexual arousal and the law of initial value: assessment at several phases of the menstrual cycle. Archives of Sexual Behavior, 22: 403413.Google Scholar
Moll, A. (1912) The Sexual Life of the Child. New York: Macmillan.Google Scholar
Montant, M., Romaiguere, P., & Roll, J. P. (2009). A new vibrator to stimulate muscle proprioceptors in fMRI. Human Brain Mapping, 30: 990997.Google Scholar
Moulier, V., Mouras, H., Pelegrini-Issac, M., Glutron, D., Rouxel, R., Grandjean, B., … & Stoléru, S. (2006). Neuroanatomical correlates of penile erection evoked by photographic stimuli in human males. NeuroImage, 33: 689699.Google Scholar
Munarriz, R., Maitland, S., Garcia, S. P., Talakoub, L., & Goldstein, I. (2003). A prospective duplex Doppler ultrasonographic study in women with sexual arousal disorder to objectively assess genital engorgement induced by EROS therapy. Journal of Sex & Marital Therapy, 29: 8594.Google Scholar
Munoz, M. M., Bancroft, J., & Marshall, I. (1993). The performance of the Rigiscan in the measurement of penile tumescence and rigidity. International Journal of Impotence Research, 5: 6976.Google Scholar
Norris, C. J., Gollan, J., Berntson, G. G., & Cacioppo, J. T. (2010). The current status of research on the structure of evaluative space. Biological Psychology, 84: 422436.Google Scholar
Oakley, S. H., Vaccaro, C. M., Crisp, C. C., Estanol, M. V., Fellner, A. N., Kleeman, S. D., & Pauls, R. N. (2014). Clitoral size and location in relation to sexual function using pelvic MRI. Journal of Sexual Medicine, 11: 10131022.Google Scholar
O’Connell, H. E., Hutson, J. M., Anderson, C. R., & Plenter, R. J. (1998). Anatomical relationship between urethra and clitoris. Journal of Urology, 159: 18921897.Google Scholar
Palace, E. M. (1995). A cognitive-physiological process model of sexual arousal and response. Clinical Psychology, Science and Practice, 2: 370384.Google Scholar
Palti, Y. & Bercovici, B. (1967). Photoplethysmographic study of the vaginal blood pulse. American Journal of Obstetrics & Gynecology, 97: 143153.Google Scholar
Payne, K. A. & Binik, Y. M. (2006). Reviving the labial thermistor clip. Archives of Sexual Behavior, 35: 111113.Google Scholar
Peterson, Z. & Janssen, E. (2007). Ambivalent affect and sexual response: the impact of co-occurring positive and negative emotions on subjective and physiological sexual responses to erotic stimuliArchives of Sexual Behavior, 36: 793807.Google Scholar
Ponseti, J. & Bosinski, H. A. (2010). Subliminal sexual stimuli facilitate genital response in women. Archives of Sexual Behavior, 39: 10731079.Google Scholar
Ponseti, J., Kropp, P., & Bosinski, H. A. (2009). Brain potentials related to the human penile erection. International Journal of Impotence Research, 21: 292300.Google Scholar
Prause, N., Barela, J., Roberts, V., & Graham, C. (2013a). Instructions to rate genital vasocongestion increases genital and self-reported sexual arousal but not coherence between genital and self-reported sexual arousal. Journal of Sexual Medicine, 10: 22192231.Google Scholar
Prause, N., Cerny, J., & Janssen, E. (2005). The labial photoplethysmograph: a new instrument for assessing genital hemodynamic changes in women. Journal of Sexual Medicine, 2: 5865.Google Scholar
Prause, N. & Harenski, C. L. (2014). Inhibition, lack of excitation or suppression: fMRI pilot of asexuality. In Cerankowski, K. J. & Milks, M. (eds.), Asexualities: Feminist and Queer Perspectives (pp. 3554). London: Taylor & Francis.Google Scholar
Prause, N. & Heiman, J. R. (2007). The labial thermistor: support for discriminant validity and utility for between-subjects designs. Psychophysiology, 44: S103.Google Scholar
Prause, N. & Heiman, J. R. (2009). Assessing female sexual arousal with the labial thermistor: response specificity and 2 construct validity. International Journal of Psychophysiology, 72: 115122.Google Scholar
Prause, N. & Heiman, J. R. (2010). Reduced labial temperature in response to sexual films with distractors among women with lower sexual desire. Journal of Sexual Medicine, 7: 951963.Google Scholar
Prause, N. & Janssen, E. (2006). Blood flow: vaginal photoplethysmography. In Goldstein, I., Meston, C. M., Davis, S., & Traish, A. (eds.), Women’s Sexual Function and Dysfunction (pp. 359367). London: Taylor & Francis.Google Scholar
Prause, N. & Rahman, A. (2015). Calibration cone for penile strain gauge [3 dimensional print file]. Retrieved from www.thingiverse.com/thing:830707Google Scholar
Prause, N., Siegle, G. J., Deblieck, C., Wu, A., & Iacoboni, M. (2016). EEG to primary rewards: predictive utility and malleability by brain stimulation. PLoS One, in press.Google Scholar
Prause, N., Staley, C., & Fong, T. W. (2013b). No evidence of emotion dysregulation in “hypersexuals” reporting their emotions to a sexual film. Sexual Addiction & Compulsivity, 20: 106126.Google Scholar
Prause, N., Staley, C., & Roberts, V. (2014a). Frontal alpha asymmetry and sexually motivated states. Psychophysiology, 51: 226235.Google Scholar
Prause, N., Steele, V. R., Staley, C., & Sabatinelli, D. (2014b). Late positive potential to explicit sexual images associated with the number of sexual intercourse partners. Social Cognitive and Affective Neuroscience, in press.Google Scholar
Prause, N., Steele, V. R., Staley, C., Sabatinelli, D., & Proudfit, G. H. (2015). Modulation of late positive potentials by sexual images in problem users and controls inconsistent with “porn addiction.” Biological Psycholology, in press.Google Scholar
Prause, N., Williams, K., & Bosworth, K. (2010). Wavelet denoising of vaginal pulse amplitude. Psychophysiology, 47: 393401.Google Scholar
Puppo, V. & Puppo, G. (2015). Anatomy of sex: revision of the new anatomical terms used for the clitoris and the female orgasm by sexologists. Clinical Anatomy, 28: 293304.Google Scholar
Rahardjo, H. E., Uckert, S., Taher, A., Sonnenberg, J. E., Kauffels, W., Rahardjo, D., & Kuczyk, M. A. (2013). Effects of endopeptidase inhibition on the contraction–relaxation response of isolated human vaginal tissue. Journal of Sexual Medicine, 10: 951959.Google Scholar
Redouté, J., Stoléru, S., Pugeat, M., Costes, N., Lavenne, F., Le Bars, D., … & Pujol, J.-F. (2005). Brain processing of visual sexual stimuli in treated and untreated hypogonadal patients. Psychoneuroendocrinology, 30: 461482.Google Scholar
Rellini, A., McCall, K. M., Meston, C. M., & Randall, P. K. (2005). The relationship between self-reported and physiological measures of female sexual arousal. Psychophysiology, 42: 116124.Google Scholar
Rellini, A. & Meston, C. (2006). The sensitivity of event logs, self-administered questionnaires and photoplethysmography to detect treatment-induced changes in female sexual arousal disorder (FSAD) diagnosis. Journal of Sexual Medicine, 3: 283291.Google Scholar
Richards, J. C., Bridger, B. A., Wood, M. M., Kalucy, R. S., & Marshall, V. R. (1985). A controlled investigation into the measurement properties of two circumferential penile strain gauges. Psychophysiology, 22: 568571.Google Scholar
Rosen, R. C. & Beck, J. G. (1988). Patterns of Sexual Arousal. New York: Guilford Press.Google Scholar
Sabatinelli, D., Bradley, M. M., Lang, P. J., Costa, V. D., & Versace, F. (2007). Pleasure rather than salience activates human nucleus accumbens and medial prefrontal cortex. Journal of Neurophysiology, 98: 13741379.Google Scholar
Salama, S., Boitrelle, F., Gauquelin, A., Malagrida, L., Thiounn, N., & Desvaux, P. (2015). Nature and origin of “squirting” in female sexuality. Journal of Sexual Medicine, 12: 661666.Google Scholar
Salonia, A., Giraldi, A., Chivers, M. L., Georgiadis, J. R., Levin, R., Maravilla, K. R., & McCarthy, M. M. (2010). Physiology of women’s sexual function: basic knowledge and new findings. Journal of Sexual Medicine, 7: 26372660.Google Scholar
Samson, A. C., Kreibig, S. D., Soderstrom, B., Wade, A. A., & Gross, J. J. (2016). Eliciting positive, negative and mixed emotional states: a film library for affective scientists. Cognition & Emotion, 30: 827856.Google Scholar
Seeley, F., Abramsen, P., Perry, L., Rothblatt, A., & Seeley, D. (1980). Thermogenic measures of sexual arousal: a methodological note. Archives of Sexual Behavior, 9: 7785.Google Scholar
Seto, M. C., Kingston, D. A., & Bourget, D. (2014). Assessment of the paraphilias. Psychiatric Clinics of North America, 37: 149161.Google Scholar
Singer, B. & Totes, F. (1987). Sexual motivation. Journal of Sex Research, 23: 481501.Google Scholar
Sintchak, G. & Geer, J. H. (1975). A vaginal plethysmograph system. Psychophysiology, 12: 113115.Google Scholar
Slob, A. K., Bax, C. M., Hop, W. C., Rowland, D. L., & van der Werff ten Bosch, J. J. (1996). Sexual arousability and the menstrual cycle. Psychoneuroendocrinology, 21: 545558.Google Scholar
Slob, A. K., Koster, J., Radder, J. K., & van der Werff ten Bosch, J. J. (1990). Sexuality and psychophysiological functioning in women with diabetes mellitus. Journal of Sex & Marital Therapy, 16: 5969.Google Scholar
Steele, V., Prause, N., Staley, C., & Fong, G. W. (2013). Sexual desire, not hypersexuality, is related to neurophysiological responses elicited by sexual images. Socioaffective Neuroscience of Psychology, 3: 20770.Google Scholar
Stoléru, S. & Mouras, H. (2007) Brain functional imaging studies of sexual desire and arousal in human males. In Janssen, E. (ed.), The Psychophysiology of Sex (pp. 334). Bloomington, IN: Indiana University Press.Google Scholar
Stoléru, S., Redoute, J., Costes, N., Lavenne, F., Bars, D., Dechaud, H., … & Pujol, J. (2003). Brain processing of visual sexual stimuli in men with hypoactive sexual desire disorder. Psychiatry Research, 124: 6786.Google Scholar
Strassberg, D. S. & Lowe, K. (1995). Volunteer bias in sexuality research. Archives of Sexual Behavior, 24: 369382.Google Scholar
Suschinsky, K. D., Bossio, J. A., & Chivers, M. L. (2014). Women’s genital sexual arousal to oral versus penetrative heterosexual sex varies with menstrual cycle phase at first exposure. Hormones and Behavior, 65: 319327.Google Scholar
Sylva, D., Safron, A., Rosenthal, A. M., Reber, P. J., Parrish, T. B., & Bailey, J. M. (2013). Neural correlates of sexual arousal in heterosexual and homosexual women and men. Hormones and Behavior, 64: 673684.Google Scholar
Tiefer, L. (1991). Historical, scientific, clinical, and feminist criticisms of “The Human Sexual Response Cycle” model. Annual Review of Sex Research, 2: 123.Google Scholar
Toates, F. (2009). An integrative theoretical framework for understanding sexual motivation, arousal, and behavior. Journal of Sex Research, 46: 168193.Google Scholar
Tucker, D. M. & Dawson, S. L. (1984). Asymmetric EEG changes as method actors generated emotions. Biological Psychology, 19: 6375.Google Scholar
Van de Velde, T. H. (1926). Ideal Marriage: Its Physiology and Technique. New York: Random House.Google Scholar
Vardi, Y., Gruenwald, I., Sprecher, E., Gertman, I., & Yartnitsky, D. (2000). Normative values for female genital sensation. Urology, 56: 10351040.Google Scholar
Vardi, Y., Sprecher, E., Gruenwald, I., Yarnitsky, D., Gartman, I., & Granovsky, Y. (2009). The P300 event-related potential technique for libido assessment in women with hypoactive sexual desire disorder. Journal of Sexual Medicine, 6: 16881695.Google Scholar
Victor, E. C., Sansosti, A. A., Bowman, H. C., & Hariri, A. R. (2015). Differential patterns of amygdala and ventral striatum activation predict gender-specific changes in sexual risk behavior. Journal of Neuroscience, 35: 88968900.Google Scholar
Voon, V., Mole, T. B., Banca, P., Porter, L., Morris, L., Mitchell, S., … & Irvine, M. (2014). Neural correlates of sexual cue reactivity in individuals with and without compulsive sexual behaviours. PLoS One, 9: e102419.Google Scholar
Voorham-van der Zalm, P. J., Lycklama à Nijeholt, G., Elzevier, H. W., Putter, H., & Pelger, R. C. (2008). Diagnostic investigation of the pelvic floor: a helpful tool in the approach in patients with complaints of micturition, defecation, and/or sexual dysfunction. Journal of Sex Research, 5: 864871.Google Scholar
Wagner, G. & Gerstenberg, T. (1988). Human in vivo studies of electrical activity of corpus cavernosum (EACC). Journal of Urology, 139: 327A.Google Scholar
Wheeler, D. & Rubin, H. B. (1987). A comparison of volumetric and circumferential measures of penile erection. Archives of Sexual Behavior, 16: 289299.Google Scholar
Wiederman, M. W. (1999). Volunteer bias in sexuality research using college student participants. Journal of Sex Research, 36: 5966.Google Scholar
Wiegel, M., Scepkowski, L., & Barlow, D. H. (2007). Cognitive-affective processes in sexual arousal and sexual dysfunction. In Janssen, E. (ed.), The Psychophysiology of Sex (pp. 143165). Bloomington, IN: Indiana University Press.Google Scholar
Wouda, J. C., Hartmen, P. M., Bakker, R. M., Bakker, J. O., van de Weil, H. B. M., Schultz, W., & Weijmar, C. M. (1998). Vaginal plethysmography in women with dyspareunia. Journal of Sex Research, 35: 141147.Google Scholar
Zuckerman, M. (1971). Physiological measures of sexual arousal in the human. Psychological Bulletin, 75: 297.Google Scholar

References

Aardal, O., Paichard, Y., Brovoll, S., Berger, T., Lande, T. S., & Hamran, S.-E. (2013). Physical working principles of medical radar. IEEE Transactions on Biomedical Engineering, 60: 11421149.Google Scholar
Aarts, L. A. M., Jeanne, V., Cleary, J. P., Lieber, C. S., Nelson, J. S., Oetomo, S. B., & Verkruysse, W. (2013). Non-contact heart rate monitoring utilizing camera photoplethysmography in the neonatal intensive care unit: a pilot study. Early Human Development, 89: 943948.Google Scholar
Adib, F., Kabelac, Z., Katabi, D., & Miller, R. C. (2014). 3D tracking via body radio reflections. Paper presented at the Usenix NSDI.Google Scholar
Adib, F., Mao, H., Kabelac, Z., Katabi, D., & Miller, R. C. (2015). Smart homes that monitor breathing and heart rate. Paper presented at the Proceedings of the 33rd Annual ACM Conference on Human Factors in Computing Systems.Google Scholar
Alametsa, J., Värri, A., Viik, J., Hyttinen, J., & Palomaki, A. (2009). Ballistocardiographic studies with acceleration and electromechanical film sensors. Medical Engineering & Physics, 31: 11541165.Google Scholar
Alekseev, S. I., Szabo, I., & Ziskin, M. C. (2008). Millimeter wave reflectivity used for measurement of skin. Skin Research and Technology, 14: 390396.Google Scholar
Alihanka, J., Vaahtoranta, K., & Saarikivi, I. (1981). A new method for long-term monitoring of the ballistocardiogram, heart rate, and respiration. American Journal of Physiology, 240: R384R392.Google Scholar
Allen, J. (2007). Photoplethysmography and its application in clinical physiological measurement. Physiological Measurement, 28: R1R39.Google Scholar
Allen, J. & Murray, A. (2003). Age-related changes in the characteristics of the photoplethysmographic pulse shape at various body sites. Physiological Measurement, 24: 297307.Google Scholar
Allen, J. A., Grimley, J. F., & Roddie, I. C. (1971). A body balance to measure sweat rates in man. Biomedical Engineering, 6: 468471.Google Scholar
Almeida, V. G., Pereira, H. C., Pereira, T., Figueiras, E., Borges, E., Cardoso, J. M. R., & Correia, C. (2011). Piezoelectric probe for pressure waveform estimation in flexible tubes and its application to the cardiovascular system. Sensors and Actuators A: Physical, 169: 217226.Google Scholar
Andersson, C., Lyass, A., Larson, M. G., Spartano, N. L., Vita, J. A., Benjamin, E. J., … & Hamburg, N. M. (2015). Physical activity measured by accelerometry and its associations with cardiac structure and vascular function in young and middle-aged adults. Journal of the American Heart Association, 4: e001528.Google Scholar
Andre, T., De Wan, M., Lefevre, P., & Thonnard, J.-L. (2008). Moisture evaluator: a direct measure of fingertip skin hydration during object manipulation. Skin Research and Technology, 14: 385389.Google Scholar
ANSI (2000). Safe Use of Lasers – ANSI Z136.1 (2000). New York: American National Standards Institute.Google Scholar
Arimoto, H. (2007). Estimation of water content distribution in the skin using dualband polarization imaging. Skin Research and Technology, 13: 4954.Google Scholar
Arnold, M. B. (1945). Physiological differentiation of emotional states. Psychological Review, 52: 3548.Google Scholar
Ax, A. F. (1953). The physiological differentiation between fear and anger in humans. Psychosomatic Medicine, 15: 433442.Google Scholar
Ayaz, H., Onaral, B., Izzetoglu, K., Shewokis, P. A., McKendrick, R., & Parasuraman, R. (2013). Continuous monitoring of brain dynamics with functional near infrared spectroscopy as a tool for neuroergonomic research: empirical examples and a technological development. Frontiers in Human Neuroscience, 7: 871.Google Scholar
Baek, H. J., Chung, G. S., Kim, K. K., Kim, J. S., & Park, K. S. (2009). Photoplethysmogram measurement without direct skin-to-sensor contact using an adaptive light source intensity control. IEEE Transactions on Information Technology in Biomedicine, 6: 10851088.Google Scholar
Baek, H. J., Kim, K. K., Kim, J. S., Lee, B., & Park, K. S. (2010). Enhancing the estimation of blood pressure using pulse arrival time and two confounding factors. Physiological Measurement, 31: 145157.Google Scholar
Baquero, G. A., Banchs, J. E., Ahmed, S., Naccarelli, G. V., & Luck, J. C. (2015). Surface 12 lead electrocardiogram recordings using smart phone technology. Journal of Electrocardiology, 48: 17.Google Scholar
Baruch, M., Kwon, K. W., Abdel-Rahman, E., & Isaacs, R. (2007). The structure of the radial pulse: a novel noninvasive ambulatory blood pressure device. In Westwood, J. D., Haluck, R. S., Hoffman, H. M., Mogel, G. T., Phillips, R., Robb, R. A., & Vosburgh, K. G. (eds.), Medicine Meets Virtual Reality 15, vol. 125 (pp. 4042). Amsterdam: IOS Press.Google Scholar
Baulmann, J., Schillings, U., Rickert, S., Uen, S., Dusing, R., Illyes, M., … & Mengden, T. (2008). A new oscillometric method for assessment of arterial stiffness: comparison with tonometric and piezo-electronic methods. Journal of Hypertension, 26: 523528.Google Scholar
Bettini, C. & Riboni, D. (2015). Privacy protection in pervasive systems: state of the art and technical challenges. Pervasive and Mobile Computing, 17: 159174.Google Scholar
Blix, A. S., Stromme, S. B., & Ursin, H. (1974). Additional heart rate: an indicator of psychological activation. Aerospace Medicine, 11: 12191222.Google Scholar
Bloch, K. E., Jugoon, S., de Socarraz, H., Manning, K., & Sackner, M. A. (1998). Thoracocardiography: noninvasive monitoring of left ventricular stroke volume. Journal of Critical Care, 13: 147157.Google Scholar
Bogler, C., Mehnert, J., Steinbrink, J., & Haynes, J.-D. (2014). Decoding vigilance with NIRS. PloS One, 9: e101729.Google Scholar
Bohannon, J. (2015). Credit card study blows holes in anonymity. Science, 347: 468.Google Scholar
Boiten, F. A., Frijda, N. H., & Wientjes, C. J. E. (1994). Emotions and respiratory patterns: a review and critical analysis. International Journal of Psychophysiology, 17: 103128.Google Scholar
Bort-Roig, J., Gilson, N. D., Puig-Ribera, A., Contreras, R. S., & Trost, S. G. (2014). Measuring and influencing physical activity with smartphone technology: a systematic review. Sports Medicine, 44: 671686.Google Scholar
Bosch, J. A., Veerman, E. C., de Geus, E. J., & Proctor, G. B. (2011). α-amylase as a reliable and convenient measure of sympathetic activity: don’t start salivating just yet! Psychoneuroendocrinology, 36: 449453.Google Scholar
Boucsein, W. (1992). Electrodermal Activity. New York: Plenum Press.Google Scholar
Boulos, M. N. K., Brewer, A. C., Karimkhani, C., Buller, D. B., & Dellavalle, R. P. (2014). Mobile medical and health apps: state of the art, concerns, regulatory control and certification. Online Journal of Public Health Informatics, 5: 229.Google Scholar
Bousefsaf, F., Maaoui, C., & Pruski, A. (2014). Remote detection of mental workload changes using cardiac parameters assessed with a low-cost webcam. Computers in Biology and Medicine, 53: 154163.Google Scholar
Branagan, M., Chenery, D. H., & Nicholson, S. (2000). Use of infrared attenuated total reflectance spectroscopy for the in vivo measurement of hydration level and silicon distribution in the stratum corneum following skin coverage by polymeric dressings. Skin Pharmacology and Applied Skin Physiology, 13: 157164.Google Scholar
Brenner, J. (1987). Behavioral energetics: some effects of uncertainty on the mobilization and distribution of energy. Psychophysiology, 24: 499512.Google Scholar
Brink, M., Muller, C., & Schierz, C. (2006). Contact-free measurement of heart rate, respiration rate, and body movements during sleep. Behavior Research Methods, 38: 511521.Google Scholar
Brovoll, S., Berger, T., Paichard, Y., Aardal, O., Lande, T. S., & Hamran, S.-E. (2014). Time-lapse imaging of human heart motion with switched array UWB radar. IEEE Transactions on Biomedical Circuits and Systems, 8: 704715.Google Scholar
Brown, H. R., Zeidman, P., Smittenaar, P., Adams, R. A., McNab, F., Rutledge, R. B., & Dolan, R. J. (2014). Crowdsourcing for cognitive science: the utility of smartphones. PloS One, 9: e100662.Google Scholar
Bruining, N., Caiani, E., Chronaki, C., Guzik, P., & van der Velde, E. (2014). Acquisition and analysis of cardiovascular signals on smartphones: potential, pitfalls and perspectives. By the Task Force of the e-Cardiology Working Group of European Society of Cardiology. European Journal of Preventive Cardiology, 21: 413.Google Scholar
Bruser, C., Stadlthanner, K., de Waele, S., & Leonhardt, S. (2011). Adaptive beat-to-beat heart rate estimation in ballistocardiograms. IEEE Transactions on Information Technology in Biomedicine, 15: 778786.Google Scholar
Buckberg, G., Hoffmann, J. I. E., Mahajan, A., Saleh, S., & Coghlan, C. (2008). Cardiac mechanics revisited: the relationship of cardiac architecture to ventricular function. Circulation, 118: 25712587.Google Scholar
Buxi, D., Redouté, J.-M., & Yuce, M. R. (2015). A survey on signals and systems in ambulatory blood pressure monitoring using pulse transit time. Physiological Measurement, 36: R1R26.Google Scholar
Cacioppo, J. T., Uchino, B. N., & Berntson, G. G. (1994). Individual differences in the autonomic origins of heart rate reactivity: the psychometrics of respiratory sinus arrhythmia and preejection period. Psychophysiology, 31: 412419.Google Scholar
Caldwell, J. A., Prazinko, B., & Caldwell, J. L. (2003). Body posture affects electroencephalographic activity and psychomotor vigilance task performance in sleep deprived subjects. Clinical Neurophysiology, 114: 2331.Google Scholar
Campbell, A., Choudhury, T., Hu, S., Lu, H., Mukerjee, M. K., Rabbi, M., & Raizada, R. D. (2010). NeuroPhone: brain–mobile phone interface using a wireless EEG headset. Paper presented at the Proceedings of the 2nd ACM SIGCOMM workshop on Networking, Systems, and Applications on Mobile Handhelds.Google Scholar
Cardone, D., Pinti, P., & Merla, A. (2015). Thermal infrared imaging-based computational psychophysiology for psychometrics. Computational and Mathematical Methods in Medicine, 2015: 984353.Google Scholar
Carroll, D., Phillips, A. C., & Balanos, G. M. (2009). Metabolically exaggerated cardiac reactions to acute psychological stress revisited. Psychophysiology, 46: 270275.Google Scholar
Casaccia, S., Sirevaag, E. J., Frank, M. G., Richter, E. J., O’Sullivan, J. A., Scalise, L., & Rohrbaugh, J. W. (2015). Noncontact sensing of facial muscle activity using laser Doppler vibrometry. Submitted.Google Scholar
Casaccia, S., Sirevaag, E. J., Richter, E. J., O’Sullivan, J. A., Scalise, L., & Rohrbaugh, J. W. (2014). Decoding carotid waveforms recorded by laser Doppler vibrometry: effects of rebreathing. Paper presented at the 11th International Conference on Vibration Measurements by Laser and Noncontact Techniques – AIVELA 2014: Advances and Applications, Ancona, Italy.Google Scholar
Case, M. A., Burwick, H. A., Volpp, K. G., & Patel, M. S. (2015). Accuracy of smartphone applications and wearable devices for tracking physical activity data. Journal of the American Medical Association, 313: 625626.Google Scholar
Castiglioni, P., Meriggi, P., Rizzo, F., Vaini, E., Faini, A., Parati, G., … & Di Rienzo, M. (2011). Cardiac sounds from a wearable device for sternal seismocardiography. Paper presented at the Engineering in Medicine and Biology Society. Annual International Conference of the IEEE.Google Scholar
Cennini, G., Arguel, J., Akşit, K., & van Leest, A. (2010). Heart rate monitoring via remote photoplethysmography with motion artifacts reduction. Optics Express, 18: 48674875.Google Scholar
Cescon, C., Farina, D., Gobbo, M., Merletti, R., & Orizio, C. (2004). Effect of accelerometer location on mechanomyogram variables during voluntary, constant-force contractions in three human muscles. Medical & Biological Engineering & Computing, 42: 121127.Google Scholar
Choi, B. & Jo, S. (2013). A low-cost EEG system-based hybrid brain–computer interface for humanoid robot navigation and recognition. PloS One, 8: e74583.Google Scholar
Clark, R. P., Goff, M. R., & MacDermot, K. D. (1990). Identification of functioning sweat pores in X-linked hypohidrotic ectodermal dysplasia by whole body thermography. Human Genetics, 86: 713.Google Scholar
Clarys, P., Clijsen, R., & Barel, A. O. (2011). Influence of probe application pressure on in vitro and in vivo capacitance (Corneometer CM 825®) and conductance (Skicon 200 EX®) measurements. Skin Research and Technology, 17: 445450.Google Scholar
Cotter, G., Schachner, A., Sasson, L., Dekel, H., & Moshkovitz, Y. (2006). Impedance cardiography revisited. Physiological Measurement, 27: 817827.Google Scholar
Coyle, S., Curto, V., Benito-Lopez, F., Florea, L., & Diamond, D. (2014). Wearable bio and chemical sensors. In Sazonoov, E. & Neuman, M. R. (eds.), Wearable Sensors: Fundamentals, Implementation and Application (pp. 6583). Amsterdam: Academic Press.Google Scholar
Curone, D., Secco, E. L., Tognetti, A., & Magenes, G. (2013). An activity classifier based on heart rate and accelerometer data fusion. International Journal of Bioelectromagnetism, 15: 712.Google Scholar
Cybulski, G. (2011). Ambulatory Impedance Cardiograph: The Sytems and their Applications. Berlin: Springer-Verlag.Google Scholar
Darrow, C. W. (1964). The rationale for treating the change in galvanic skin response as a change in conductance. Psychophysiology, 1: 3138.Google Scholar
de Groot, J. H. B., Smeets, M. A. M., Rowson, M. J., Bulsing, P. J., Blonk, C. G., Wilkinson, J. E., & Semin, G. R. (2015). A sniff of happiness. Psychological Science, 26: 684700.Google Scholar
de Haan, G. & van Leest, A. (2014). Improved motion robustness of remote-PPG by using the blood volume pulse signature. Physiological Measurement, 35: 19131926.Google Scholar
De Melis, M., Morbiducci, U., & Scalise, L. (2007). Identification of cardiac events by optical vibrocardiography: comparison with phonocardiography. Paper presented at the 29th Annual International Conference of the IEEE Engineering and Biomedicine Society, Lyon, France, August 23–26.Google Scholar
Dehkordi, P., Marzencki, M., Tavakolian, K., Kaminska, M., & Kaminska, B. (2012). Monitoring torso acceleration for estimating the respiratory flow and efforts for sleep apnea detection. Paper presented at the Engineering in Medicine and Biology Society. Annual International Conference of the IEEE.Google Scholar
Di Rienzo, M., Meriggi, P., Rizzo, F., Vaini, E., Faini, A., Merati, G., … & Castiglioni, P. (2011). A wearable system for the seismocardiogram assessment in daily life conditions. Paper presented at the Engineering in Medicine and Biology Society. Annual International Conference of the IEEE.Google Scholar
Di Rienzo, M., Vaini, E., Castiglioni, P., Merati, G., Meriggi, P., Parati, G., … & Rizzo, F. (2013). Wearable seismocardiography: towards a beat-by-beat assessment of cardiac mechanics in ambulant subjects. Autonomic Neuroscience, 178: 5059.Google Scholar
Dittmar, A., Gehin, C., Delhomme, G., Boivin, D., Dumont, G., & Mott, C. (2006). A noninvasive wearable sensor for the measurement of brain temperature. Paper presented at the Engineering in Medicine and Biology Society, 28th Annual International Conference of the IEEE.Google Scholar
Dräbenstedt, A., Sauer, J., & Rembe, C. (2012). Remote-sensing vibrometry at 1550 nm wavelength. Paper presented at the 10th International Conference on Vibration Measurements by Laser and Noncontact Techniques – AIVELA, AIP Conference Proceedings.Google Scholar
Drysdale, D. (2014). Transcutaneous carbon dioxide monitoring: literature review. Oral Health and Dental Management, 13: 453457.Google Scholar
Duschek, S. & Schandry, R. (2003). Functional transcranial Doppler sonography as a tool in psychophysiological research. Psychophysiology, 40: 436454.Google Scholar
Duvinage, M., Castermans, T., Dutoit, T., Petieau, M., Hoellinger, T., De Saedeleer, C., … & Cheron, G. (2012). A P300-based quantitative comparison between the Emotiv Epoc headset and a medical EEG device. Biomedical Engineering, 765: 20122764.Google Scholar
Eddleman, E. E. J. (1974). Kinetocardiography. In Weissler, A. M. (ed.), Noninvasive Cardiology (pp. 227273). New York: Grune & Stratton.Google Scholar
Edelberg, R. & Wright, D. J. (1964). Two galvanic skin response effector organs and their stimulus specificity. Psychophysiology, 1: 3947.Google Scholar
Edens, J. L., Larkin, K. T., & Abel, J. L. (1992). The effect of social support and physical touch on cardiovascular reactions to mental stress. Journal of Psychosomatic Research, 36: 371382.Google Scholar
Ekandem, J. I., Davis, T. A., Alvarez, I., James, M. T., & Gilbert, J. E. (2012). Evaluating the ergonomics of BCI devices for research and experimentation. Ergonomics, 55: 592598.Google Scholar
Ekman, P., Friesen, W. V., & Hager, J. C. (2002). Facial Action Coding System: The Manual. Salt Lake City, UT: Research Nexus.Google Scholar
Ernst, J. M., Litvack, D. A., Lozano, D. L., Cacioppo, J. T., & Berntson, G. G. (1999). Impedance pneumography: noise as signal in impedance cardiography. Psychophysiology, 36: 333338.Google Scholar
Etemadi, M., Inan, O. T., Heller, J. A., Hersek, S., Klein, L., & Roy, S. (2016). A wearable patch to enable long-term monitoring of environmental, activity and hemodynamics variables. IEEE Transactions on Biomedical Circuits and Systems, 10: 8088.Google Scholar
Fei, J. & Pavlidis, I. (2010). Thermistor at a distance: unobtrusive measurement of breathing. IEEE Transactions on Biomedical Engineering, 57: 988998.Google Scholar
Fei, J., Zhu, Z., & Pavlidis, I. (2005). Imaging breathing rate in the CO2 absorption band. Paper presented at the Engineering in Medicine and Biology Society, 27th Annual International Conference of the IEEE.Google Scholar
Feldman, Y., Puzenko, A., Ben Ishai, P., Caduff, A., & Agranat, A. J. (2008). Human skin as arrays of helical antennas in the millimeter and submillimeter wave range. Physical Review Letters, 100: 128102.Google Scholar
Ferscha, A. (2014). Attention, please! IEEE Pervasive Computing, 13: 4854.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. 5196). New York: Guilford Press.Google Scholar
Freedman, L. W., Scerbo, A. S., Dawson, M. E., Raine, A., McClure, W. O., & Venables, P. H. (1994). The relationship of sweat gland count to electrodermal activity. Psychophysiology, 31: 186200.Google Scholar
Fujikawa, T., Tochikubo, O., Kura, N., Kiyokura, T., Shimada, J., & Umemura, S. (2009). Measurement of hemodynamics during postural changes using a new wearable cephalic laser blood flowmeter. Circulation Journal, 73: 19501955.Google Scholar
Gale, A. & Baker, S. (1981). In vivo or in vitro? Some effects of laboratory environments, with particular reference to the psychophysiology experiment. In Christie, M. J. & Mellett, P. G. (eds.), Foundations of Psychosomatics (pp. 363384). London: John Wiley.Google Scholar
Gallace, A. & Spence, C. (2010). The science of interpersonal touch: an overview. Neuroscience & Biobehavioral Reviews, 34: 246259.Google Scholar
Gandhi, N., Khe, C., Chung, D., Chi, Y. M., & Cauwenberghs, G. (2011). Properties of dry and non-contact electrodes for wearable physiological sensors. Paper presented at the International Conference on Body Sensor Networks (BSN).Google Scholar
Garbey, M., Merla, A., & Pavlidis, I. T. (2004). Estimation of blood flow speed and vessel location from thermal video. Paper presented at the IEEE Conference on Computer Vision and Pattern Recognition.Google Scholar
Gesche, H., Grosskurth, D., Küchler, G., & Patzak, A. (2012). Continuous blood pressure measurement by using the pulse transit time: comparison to a cuff-based method. European Journal of Applied Physiology, 112: 309315.Google Scholar
Giannattasio, C., Vincenti, A., Failla, M., Capra, A., Ciro, A., De Ceglia, S., … & Mancia, G. (2003). Effects of heart rate changes on arterial distensibility in humans. Hypertension, 42: 253256.Google Scholar
Giovangrandi, L., Inan, O. T., Wiard, R. M., Etemadi, M., & Kovacs, G. T. A. (2011). Ballistocardiography: a method worth revisiting. Paper presented at the 33rd Annual International Conference of the IEEE EMBS, Boston, MA.Google Scholar
Goodwin, M. S., Velicer, W. F., & Intille, S. S. (2008). Telemetric monitoring in the behavior sciences. Behavior Research Methods, 40: 328341.Google Scholar
Gosse, P., Guillo, P., Ascher, G., & Clementy, J. (1994). Assessment of arterial distensibility by monitoring the timing of Korotkoff sounds. American Journal of Hypertension, 7: 228233.Google Scholar
Grabell, M. & Salewski, C. (2011). Sweating bullets: body scanners can see perspiration as a potential weapon. ProPublica, December 19. Retrieved from www.propublica.org/article/sweating-bullets-body-scanners-can-see-perspiration-as-a-potential-weapon,22900.Google Scholar
Gramann, K., Gwin, J. T., Ferris, D. P., Oie, K., Jung, T.-P., Lin, C.-T., … & Makeig, S. (2011). Cognition in action: imaging brain/body dynamics in mobile humans. Reviews in the Neurosciences, 22: 593608.Google Scholar
Greneker, E. F. (1997). Radar sensing of heartbeat and respiration at a distance with applications of the technology. Paper presented at the Radar 97 Conference.Google Scholar
Gribok, A., Rumpler, W., Hines, W., Hoyt, R., & Buller, M. (2014). Subcutaneous glucose concentration as a predictor variable for energy expenditure during resistance exercise in humans. Paper presented at the 11th International Conference on Wearable and Implantable Body Sensor Networks (BSN).Google Scholar
Gurovich, A. N., Beck, D. T., & Braith, R. W. (2009). Aortic pulse wave analysis is not a surrogate for central arterial pulse wave velocity. Experimental Biology and Medicine, 234: 13391344.Google Scholar
Hagenaars, M. A., Oitzl, M., & Roelofs, K. (2014). Updating freeze: aligning animal and human research. Neuroscience & Biobehavioral Reviews, 47, 165176.Google Scholar
Hasegawa, M., Rodbard, D., & Kinoshita, Y. (1991). Timing of the carotid arterial sounds in normal adult men: measurement of left ventricular ejection, pre-ejection period and pulse transmission time. Cardiology, 78: 138149.Google Scholar
Hayashi, N., Someya, N., Maruyama, T., Hirooka, Y., Endo, M. Y., & Fukuba, Y. (2009). Vascular responses to fear-induced stress in humans. Physiology & Behavior, 98: 441446.Google Scholar
He, D. D., Winokur, E. S., & Sodini, C. G. (2011). A continuous, wearable, and wireless heart monitor using head ballistogram (BCG) and head electrocardiogram (ECG). Paper presented at the 33rd Annual International Conference of the IEEE EMBS, Boston, MA.Google Scholar
Head, G. A. (2014). Ambulatory blood pressure is ready to replace clinic blood pressure in the diagnosis of hypertension: pro side of the argument. Hypertension, 64: 11751181.Google Scholar
Heffernan, K. S., Spartano, N. L., Augustine, J. A., Lefferts, W. K., Hughes, W. E., Mitchell, G. F., … & Gump, B. B. (2015). Carotid artery stiffness and hemodynamic pulsatility during cognitive engagement in healthy adults: a pilot investigation. American Journal of Hypertension, 28: 615622.Google Scholar
Hennig, J., Friebe, J., Ryl, I., Kramer, B., Bottcher, J., & Netter, P. (2000). Upright posture influences salivary cortisol. Psychoneuroendocrinology, 25: 6983.Google Scholar
Heo, Y. J. & Takeuchi, S. (2013). Towards smart tattoos: implantable biosensors for continuous glucose monitoring. Advanced Healthcare Materials, 2: 4356.Google Scholar
Hertenstein, M. J., Verkamp, J. M., Kerestes, A. M., & Holmes, R. M. (2006). The communicative functions of touch in humans, nonhuman primates, and rats: a review and synthesis of the empirical research. Genetic, Social, and General Psychology Monographs, 132: 594.Google Scholar
Hirabayashi, M., Fujiwara, C., Ohtani, N., Kagawa, S., & Kamide, M. (2009). Transcutaneous PCO2 monitors are more accurate than end-tidal PCO2 monitors. Journal of Anesthesia, 23: 198202.Google Scholar
Hirschberg, D. L., Betts, K., Emanuel, P., & Caples, M. (2014). Assessment of Wearable Sensor Technologies for Biosurveillance (ECBC-TR-1275). Aberdeen, MD: Edgewood Chemical Biological Center.Google Scholar
Holland, C. & Komogortsev, O. (2012). Eye tracking on unmodified common tablets: challenges and solutions. Paper presented at the Proceedings of the Symposium on Eye Tracking Research and Applications.Google Scholar
Holter, N. J. (1961). New method for heart studies. Science, 134: 12141220.Google Scholar
Hope, S. A., Meredith, I. T., & Cameron, J. D. (2004). Effect of non-invasive calibration of radial waveforms on error in transfer-function-derived central aortic waveform characteristics. Clinical Science, 107: 205211.Google Scholar
Houtveen, J. H. & de Geus, E. J. C. (2009). Noninvasive psychophysiological ambulatory recordings: study design and data analysis strategies. European Psychologist, 14: 132141.Google Scholar
Houtveen, J. H., Hamaker, E. L., & Van Doornen, L. J. P. (2010). Using multilevel path analysis in analyzing 24-h ambulatory physiological recordings applied to medically unexplained symptoms. Psychophysiology, 47: 570578.Google Scholar
Hu, Y., Kim, E. G., Cao, G., Liu, S., & Xu, Y. (2014). Physiological acoustic sensing based on accelerometers: a survey for mobile healthcare. Annals of Biomedical Engineering, 42: 22642277.Google Scholar
Huis In ’t Veld, E. M. J., van Boxtel, G. J. M., & de Gelder, B. (2014). The Body Action Coding System II: muscle activations during the perception and expression of emotion. Frontiers in Behavioral Neuroscience, 8: 330.Google Scholar
Humeau-Heurtier, A., Guerreschi, E., Abraham, P., & Mahe, G. (2013). Relevance of laser Doppler and laser speckle techniques for assessing vascular function: state of the art and future trends. IEEE Transactions on Biomedical Engineering, 60: 659666.Google Scholar
Inan, O. T., Etemadi, M., Wiard, R. M., Giovangrandi, L., & Kovacs, G. T. A. (2009). Robust ballistocardiogram acquisition for home monitoring. Physiological Measurement, 30: 169185.Google Scholar
Intille, S. S. (2012). Emerging technology for studying daily life. In Mehl, M. R. & Conner, T. S. (eds.), Handbook of Research Methods for Studying Daily Life (pp. 267282). New York: Guilford Press.Google Scholar
Ioannou, S., Gallese, V., & Merla, A. (2014). Thermal infrared imaging in psychophysiology: potentialities and limits. Psychophysiology, 51: 951963.Google Scholar
Ionescu, V., Tarlea, M., Palaghita, M., & Moraru, N. (1985). Phonocardiographic changes induced by neuropsychic stress. Physiologie, 22: 249255.Google Scholar
Irani, R., Nasrollahi, K., & Moeslund, T. B. (2014). Improved pulse detection from head motions using DCT. Paper presented at the 9th International Conference on Computer Vision Theory and Applications.Google Scholar
Jain, A., Schmidt, T. F. H., Johnston, D. W., Brabant, G., & von zur Mühlen, A. (1998). The relationship between heart rate and blood pressure reactivity in the laboratory and in the field: evidence using continuous measures of blood pressure, heart rate and physical activity. Journal of Psychophysiology, 12: 362375.Google Scholar
Jakovljevic, D. G., Moore, S., Hallsworth, K., Fattakhova, G., Thoma, C., & Trenell, M. I. (2012). Comparison of cardiac output determined by bioimpedance and bioreactance methods at rest and during exercise. Journal of Clinical Monitoring and Computing, 26: 6368.Google Scholar
Jatoi, N.-A., Kyvelou, S.-M., & Feely, J. (2014). The acute effects of mental arithmetic, cold pressor and maximal voluntary contraction on arterial stiffness in young healthy subjects. Artery Research, 8: 4450.Google Scholar
Jean-Louis, G., Kripke, D. F., Mason, W. J., Elliott, J. A., & Youngstedt, S. D. (2001). Sleep estimation from wrist movement quantified by different actigraphic modalities. Journal of Neuroscience Methods, 105: 185191.Google Scholar
Jeanne, V., Asselman, M., den Brinker, B., & Bulut, M. (2013). Camera-based heart rate monitoring in highly dynamic light conditions. Paper presented at the International Conference on Connected Vehicles and Expo (ICCVE).Google Scholar
Jerrard-Dunne, P., Mahmud, A., & Feely, J. (2008). Ambulatory stiffness index, pulse wave velocity and augmentation index: interchangeable or mutually exclusive measures? Journal of Hypertension, 26: 529534.Google Scholar
Johnson, C. & Shuster, S. (1969). The measurement of transepidermal water loss. British Journal of Dermatology, 81: 4046.Google Scholar
Jones, A. Y. M. & Dean, E. (2004). Body position change and its effects on hemodynamic and metabolic status. Heart & Lung, 33: 281290.Google Scholar
Juen, J., Cheng, Q., Prieto-Centurion, V., Krishnan, J. A., & Schatz, B. (2014). Health monitors for chronic disease by gait analysis with mobile phones. Telemedicine and e-Health, 20: 10351041.Google Scholar
Junnila, S., Akhbardeh, A., & Värri, A. (2009). An electromechanical film sensor based wireless ballistocardiographic chair: implementation and performance. Journal of Signal Processing and Systems, 57: 305320.Google Scholar
Kamshilin, A. A., Miridonov, S., Teplov, V., Saarenheimo, R., & Nippolainen, E. (2011). Photoplethysmographic imaging of high spatial resolution. Biomedical Optics Express, 2: 9961006.Google Scholar
Kaniusas, E., Pfutzner, H., Mehnen, L., Kosel, J., Varoneckas, G., Alonderis, A., & Zakarevicius, L. (2008). Cardiovascular oscillations of the carotid artery assessed by magnetoelastic skin curvature sensor. IEEE Transactions on Biomedical Engineering, 55: 369372.Google Scholar
Kaplan, R. M. & Stone, A. A. (2013). Bringing the laboratory and clinic to the community: mobile technologies for health promotion and disease prevention. Annual Review of Psychology, 64: 471498.Google Scholar
Kashima, H., Hamada, Y., & Hayashi, N. (2014). Palatability of tastes is associated with facial circulatory responses. Chemical Senses, 39: 243248.Google Scholar
Katranas, G. S., Meydan, T., Ovari, T.-A., & Borza, F. (2008). Applications of the bi-layer thin film sensor system for registering cardio-respiratory activity. Sensors and Actuators A: Physical, 142: 455458.Google Scholar
Kaur, B., Hutchinson, J. A., Leonard, K. R., & Nelson, J. K. (2011). Human facial skin detection in thermal video to effectively measure electrodermal activity (EDA). Paper presented at the SPIE Defense, Security, and Sensing Conference.Google Scholar
Kelly, R., Karamanoglu, M., Gibbs, H. H., Avolio, A. P., & O’Rourke, M. F. (1989). Noninvasive carotid pressure wave registration as an indicator of ascending aortic pressure. Journal of Vascular Medicine and Biology, 1: 241247.Google Scholar
Kerassidis, S. (1994). Is palmar and plantar sweating thermoregulatory? Acta Physiologia Scandinavica, 152: 259263.Google Scholar
Khan, W. Z., Xiang, Y., Aalsalem, M. Y., & Arshad, Q. (2013). Mobile phone sensing systems: a survey. IEEE Communications Surveys & Tutorials, 15(1): 402427.Google Scholar
Kim, D.-H., Lu, N., Ma, R., Kim, Y.-S., Kim, R.-H., Wang, S., … & Islam, A. (2011). Epidermal electronics. Science, 333: 838843.Google Scholar
Kim, J. S., Chee, Y. J., Park, J. W., Choi, J. W., & Park, K. S. (2006). A new approach for non-intrusive monitoring of blood pressure on a toilet seat. Physiological Measurement, 27: 203213.Google Scholar
Kim, K.-K., Chee, Y.-J., Lim, Y.-G., Choi, J.-W., & Park, K.-S. (2006). A new method for unconstrained pulse arrival time (PAT) measurement on a chair. Journal of Biomedical Engineering Research, 27: 8388.Google Scholar
Kohler, T. & Schuschel, I. (1994). Changes in the number of active sweat glands (palmar sweat index, PSI) during a distressing film. Biological Psychology, 37: 133145.Google Scholar
Krishna, S., Boren, S. A., & Balas, E. A. (2009). Healthcare via cell phones: a systematic review. Telemedicine and e-Health, 15: 231240.Google Scholar
Krzywicki, A. T., Berntson, G. G., & O’Kane, B. L. (2014). A non-contact technique for measuring eccrine sweat gland activity using passive thermal imaging. International Journal of Psychophysiology, 94: 2534.Google Scholar
Kudo, H. & Mitsubayashi, K. (2012). Flexible and wearable chemical sensors for noninvasive monitoring. In Lai, D. T. H., Begg, R., & Palaniswami, M. (eds.), Healthcare Sensor Networks: Challenges Toward Practical Implementation (pp. 139157). Boca Raton, FL: CRC Press.Google Scholar
Kuipers, N. T., Sauder, C. L., Carter, J. R., & Ray, C. A. (2008). Neurovascular responses to mental stress in the supine and upright postures. Journal of Applied Physiology, 104: 11291136.Google Scholar
Kundi, M. (2009). The controversy about a possible relationship between mobile phone use and cancer. Environmental Health Perspectives, 117: 316324.Google Scholar
Kuno, Y. (1956). Human Perspiration. Springfield, IL: Charles C. Thomas.Google Scholar
Kurylyak, Y., Lamonaca, F., & Grimaldi, D. (2013). A neural network-based method for continuous blood pressure estimation from a PPG signal. Paper presented at the IEEE International Conference on Instrumentation and Measurement Technology (I2MTC).Google Scholar
Labrador, M. A. & Lara Yejas, O. D. (2014). Human Activity Recognition Using Wearable Sensors and Smartphones. Boca Raton, FL: CRC Press.Google Scholar
Lakens, D. (2013). Using a smartphone to measure heart rate changes during relived happiness and anger. IEEE Transactions on Affective Computing, 4: 238241.Google Scholar
Lamonaca, F., Barbe, K., Kurylyak, Y., Grimaldi, D., Van Moer, W., Furfaro, A., & Spagnuolo, V. (2013). Application of the Artificial Neural Network for blood pressure evaluation with smartphones. Paper presented at the IEEE 7th International Conference on Intelligent Data Acquisition and Advanced Computing Systems (IDAACS).Google Scholar
Lane, J. D., Greenstadt, L., Shapiro, D., & Rubenstein, E. (1983). Pulse transit time and blood pressure: an intensive analysis. Psychophysiology, 20: 4549.Google Scholar
Lee, Y.-D. & Chung, W.-Y. (2009). Wireless sensor network based wearable smart shirt for ubiquitous health and activity monitoring. Sensors and Actuators B: Chemical, 140: 390395.Google Scholar
Leite, I., Martinho, C., & Paiva, A. (2013). Social robots for long-term interaction: a survey. International Journal of Social Robotics, 5: 291308.Google Scholar
Lemay, M., Mertschi, M., Sola, J., Renevey, P., Parak, J., & Korhonen, I. (2014). Application of optical heart rate monitoring. In Sazonov, E. & Neuman, M. R. (eds.), Wearable Sensors: Fundamentals, Implemenation and Applications (pp. 105129). Amsterdam: Academic Press.Google Scholar
Lewis, G. F., Gatto, R., & Porges, S. W. (2011). A novel method for extracting respiration rate and relative tidal volume from infrared thermography. Psychophysiology, 48: 877887.Google Scholar
Lim, Y. G., Hong, K. H., Kim, K. K., Shin, J. H., Lee, S. M., Chung, G. S., … & Park, K. S. (2011). Monitoring physiological signals using nonintrusive sensors installed in daily life equipment. Biomedical Engineering Letters, 1: 1120.Google Scholar
Lin, J. & Wu, W. (2014). Vital sign radars: past, present, and future. Paper presented at the IEEE 15th Annual Conference on Wireless and Microwave Technology (WAMICON).Google Scholar
Liu, H., Ivanov, K., Wang, Y., & Wang, L. (2015). A novel method based on two cameras for accurate estimation of arterial oxygen saturation. Biomedical Engineering Online, 14: 52.Google Scholar
Liu, J., Wang, Y., Chen, Y., Yang, J., Chen, X., & Cheng, J. (2015). Tracking vital signs during sleep leveraging off-the-shelf WiFi. Paper presented at the Proceedings of the 16th ACM International Symposium on Mobile Ad Hoc Networking and Computing.Google Scholar
Logan, A. G. (2013). Transforming hypertension management using mobile health technology for telemonitoring and self-care support. Canadian Journal of Cardiology, 29: 579585.Google Scholar
Lovett, P. B., Buchwald, J. M., Sturmann, K., & Bijur, P. (2005). The vexatious vital: neither clinical measurements by nurses nor an electronic monitor provides accurate measurements of respiratory rate of triage. Annals of Emergency Medicine, 45: 6876.Google Scholar
Lu, L., Li, C., & Lie, D. Y. (2010). Experimental demonstration of noncontact pulse wave velocity monitoring using multiple Doppler radar sensors. Paper presented at the Engineering in Medicine and Biology Society, Annual International Conference of the IEEE.Google Scholar
Luisada, A. A., MacCanon, D. M., Coleman, B., & Feigen, L. P. (1971). New studies on the first heart sound. American Journal of Cardiology, 28: 140149.Google Scholar
Machado-Moreira, C. A., Barry, R. J., Vosselman, M. J., Ruest, R. M., & Taylor, N. A. S. (2015). Temporal and thermal variations in site-specific thermoregulatory sudomotor thresholds: precursor versus discharged sweat production. Psychophysiology, 52: 117123.Google Scholar
MacKerron, G. & Mourato, S. (2013). Happiness is greater in natural environments. Global Environmental Change, 23: 9921000.Google Scholar
Mackinnon, A. D., Aaslid, R., & Markus, H. S. (2004). Long-term ambulatory monitoring for cerebral emboli using transcranial Doppler ultrasound. Stroke, 35: 7378.Google Scholar
Makikawa, M., Shiozawa, N., & Okada, S. (2014). Fundamentals of wearable sensors for the monitoring of physical and physiological changes in daily life. In Sazonov, E. & Neuman, M. R. (eds.), Wearable Sensors: Fundamentals, Implementation and Applications (pp. 517541). Amsterdam: Academic Press.Google Scholar
Marciano, F., Cammarota, S., Migaux, M. L., Ferro, G., & Rentsch, W. (1992). Pulse transducers for long term measurement of systolic time intervals: applications to Holter recordings. Paper presented at the IEEE Engineering in Medicine and Biology Conference.Google Scholar
Mateu, L., Drager, T., Mayordomo, I., & Pollak, M. (2014). Energy harvesting at the human body. In Sazonov, E. & Neuman, M. R. (eds.), Wearable Sensors: Fundamentals, Implementation and Applications (pp. 235298). Amsterdam: Academic Press.Google Scholar
Maton, B., Petitjean, M., & Cnockaert, J. C. (1990). Phonomyogram and electromyogram relationships with isometric force reinvestigated in man. European Journal of Applied Physiology, 60: 194201.Google Scholar
Matzeu, G., Florea, L., & Diamond, D. (2015). Advances in wearable chemical sensor design for monitoring biological fluids. Sensors and Actuators B: Chemical, 211: 403418.Google Scholar
McCaffrey, C., Chevalerias, O., Mathuna, C. O., & Twomey, K. (2008). Swallowable-capsule technology. IEEE Pervasive Computing, 7: 2329.Google Scholar
McDuff, D., Gontarek, S., & Picard, R. W. (2014a). Improvements in remote cardiopulmonary measurement using a five band digital camera. IEEE Transactions on Biomedical Engineering, 61: 25932601.Google Scholar
McDuff, D., Gontarek, S., & Picard, R. (2014b). Remote measurement of cognitive stress via heart rate variability. Paper presented at the Engineering in Medicine and Biology Society, 36th Annual International Conference of the IEEE.Google Scholar
McGregor, I. A. (1952). The sweating reactions of the forehead. Journal of Physiology, 116: 2634.Google Scholar
McKay, W. P., Gregson, P. H., McKay, B. W., & Militzer, J. (1999). Sternal acceleration ballistocardiography and arterial pressure wave analysis to determine stroke volume. Clinical and Investigative Medicine, 22: 414.Google Scholar
McKendrick, R., Ayaz, H., Olmstead, R., & Parasuraman, R. (2014). Enhancing dual-task performance with verbal and spatial working memory training: continuous monitoring of cerebral hemodynamics with NIRS. NeuroImage, 85: 10141026.Google Scholar
McKendrick, R., Parasuraman, R., & Ayaz, H. (2015). Wearable functional near infrared spectroscopy (fNIRS) and transcranial direct current stimulation (tDCS): expanding vistas for neurocognitive augmentation. Frontiers in Systems Neuroscience, 9: 27.Google Scholar
McLaughlin, J., McNeill, M., Braun, B., & McCormack, P. D. (2003). Piezoelectric sensor determination of arterial pulse wave velocity. Physiological Measurement, 24: 693702.Google Scholar
McNair, D. M., Droppleman, L. F., & Pillard, R. C. (1967). Differential sensitivity of two palmar sweat measures. Psychophysiology, 3: 280284.Google Scholar
Mehl, M. R. & Conner, T. S. (2012). Handbook of Research Methods for Studying Daily Life. New York: Guilford Press.Google Scholar
Merla, A., Di Donato, L., Romani, G. L., & Rossini, P. M. (2003). Recording of the sympathetic thermal response by means of infrared functional imaging. Paper presented at the Engineering in Medicine and Biology, 25th Annual International Conference of the IEEE, Cancun, Mexico.Google Scholar
Merla, A., Di Romualdo, L., Proietti, M., Salsano, F., & Romani, G. L. (2007). Combined thermal and laser Doppler imaging in the assessment of cutaneous tissue perfusion. Paper presented at the 29th Annual International Conference of the IEEE EMBS, Lyon, France, August 23–26.Google Scholar
Meyer, T., Davison, R. C., & Kindermann, W. (2005). Ambulatory gas exchange measurements: current status and future options. International Journal of Sports Medicine, 26: S19S27.Google Scholar
Millasseau, S. & Agnoletti, D. (2015). Non-invasive estimation of aortic blood pressures: a close look at current devices and methods. Current Pharmaceutical Design, 21: 709718.Google Scholar
Miller, G. (2012). The smartphone psychology manifesto. Perspectives on Psychological Science, 7: 221237.Google Scholar
Mitchem, J. C. & Tuttle, W. W. (1954). Influence of exercises, emotional stress, and age on static neuromuscular tremor magnitude. Research Quarterly of the American Association for Health, Physical Education, & Recreation, 25: 6574.Google Scholar
Miyamoto, N. & Oda, S. (2003). Mechanomyographic and electromyographic responses of the triceps surae during maximal voluntary contractions. Journal of Electromyography and Kinesiology, 13: 451459.Google Scholar
Morrison, I., Löken, L. S., & Olausson, H. (2010). The skin as a social organ. Experimental Brain Research, 204: 305314.Google Scholar
Muehlhan, M., Marxen, M., Landsiedel, J., Malberg, H., & Zaunseder, S. (2014). The effect of body posture on cognitive performance: a question of sleep quality. Frontiers in Human Neuroscience, 8: 171.Google Scholar
Muiesan, M. L., Salvetti, M., Bertacchini, F., Agabiti-Rosei, C., Mauruelli, G., & Colonetti, E. (2014). Central blood pressure assessment using 24-hour brachial pulse wave analysis. Journal of Vascular Diagnostics, 2: 141148.Google Scholar
Murthy, J. N., van Jaarsveld, J., Fei, J., Pavlidis, I., Harrykissoon, R. I., Lucke, J. F., … & Castriotta, R. J. (2009). Thermal infrared imaging: a novel method to monitor airflow during polysomnography. Sleep, 32: 15211527.Google Scholar
Nakajima, K., Maekawa, T., & Miike, H. (1997). Detection of apparent skin motion using optical flow analysis: blood pulsation signal obtained from optical flow sequence. Review of Scientific Instruments, 68: 13311336.Google Scholar
Nam, Y., Lee, J., & Chon, K. H. (2014). Respiratory rate estimation from the built-in cameras of smartphones and tablets. Annals of Biomedical Engineering, 42: 885898.Google Scholar
Narayan, O., Casan, J., Szarski, M., Dart, A. M., Meredith, I. T., & Cameron, J. D. (2014). Estimation of central aortic blood pressure: a systematic meta-analysis of available techniques. Journal of Hypertension, 32: 17271740.Google Scholar
Ng, K.-G. (2011). Review of measurement methods and clinical validation studies of noninvasive blood pressure monitors: accuracy requirements and protocol considerations for devices that require patient-specific calibration by a secondary method or device before use. Blood Pressure Monitoring, 16: 291303.Google Scholar
Ng, K.-G., Ting, C.-M., Yeo, J.-H., Sim, K.-W., Peh, W.-L., Chua, N.-H., … & Kwong, F. (2004). Progress on the development of the MediWatch ambulatory blood pressure monitor and related devices. Blood Pressure Monitoring, 9: 149165.Google Scholar
Nichols, W. W., O’Rourke, M. F., & Vlachopoulos, C. (2011). McDonald’s Blood Flow in Arteries: Theoretical, Experimental and Clinical Principles, 6th edn. New York: CRC Press.Google Scholar
Nilsen, W. J. & Vrana, S. R. (1998). Some touching situations: the relationship between gender and contextual variables in cardiovascular responses to human touch. Annals of Behavioral Medicine, 20: 270276.Google Scholar
Nilsson, L. (1982). Relationship between evoked skin-conductance response and evaporative-water-loss response following acoustic stimulation. Medical & Biomedical Engineering & Computing, 20: 687692.Google Scholar
Nilsson, L. M. (2013). Respiration signals from photoplethysmography. Anesthesia & Analgesia, 117: 859865.Google Scholar
Obrist, P. A., Light, K. C., McCubbin, J. A., Hutcheson, J. S., & Hoffer, J. L. (1978). Pulse transit time: relationship to blood pressure. Behavior Research Methods and Instrumentation, 10: 623626.Google Scholar
Odinaka, I., Lai, P.-H., Kaplan, A. D., O’Sullivan, J. A., Sirevaag, E. J., & Rohrbaugh, J. W. (2012). ECG biometric recognition: a comparative analysis. IEEE Transactions on Information Forensics and Security, 7: 18121824.Google Scholar
Ogata, A., Sugenoya, J., Nishimura, N., & Matsumoto, T. (2005). Low and high frequency acupuncture stimulation inhibits mental stress-induced sweating in humans via different mechanisms. Autonomic Neuroscience: Basic and Clinical, 118: 93101.Google Scholar
Ohhashi, T., Sakaguchi, M., & Tsuda, T. (1998). Human perspiration measurement. Physiological Measurement, 19: 449461.Google Scholar
Ohkubo, T., Imai, Y., Tsuji, I., Nagai, K., Kato, J., Kikuchi, N., … & Kikuya, M. (1998). Home blood pressure measurement has a stronger predictive power for mortality than does screening blood pressure measurement: a population-based observation in Ohasama, Japan. Journal of Hypertension, 16: 971975.Google Scholar
Omboni, S., Gazzola, T., Carabelli, G., & Parati, G. (2013). Clinical usefulness and cost effectiveness of home blood pressure telemonitoring: meta-analysis of randomized controlled studies. Journal of Hypertension, 31: 455468.Google Scholar
Orizio, C., Gobbo, M., Diemont, B., Esposito, F., & Veicsteinas, A. (2003a). The surface mechanomyogram as a tool to describe the infuence of fatigue on biceps brachii motor unit activation strategy: historical basis and novel evidence. European Journal of Applied Physiology, 90: 326336.Google Scholar
Orizio, C., Gobbo, M., Veicsteinas, A., Baratta, R. V., Zhou, B. H., & Solomonow, M. (2003b). Transients of the force and surface mechanomyogram during cat gastrocnemius tetanic stimulation. European Journal of Applied Physiology, 88: 601606.Google Scholar
Ouwerkerk, M., Dandine, P., Bolio, D., Kocielnik, R., Mercurio, J., Huijgen, H., & Westerink, J. (2013). Wireless multi sensor bracelet with discreet feedback. Paper presented at the Proceedings of the 4th Conference on Wireless Health.Google Scholar
Paalasmaa, J. (2010). A respiratory latent variable model for mechanically measured heartbeats. Physiological Measurement, 31: 13311344.Google Scholar
Pakrashi, V. & Wirth, G. (2013). Enhanced laser Doppler vibrometer technology enables non-contact vibration measurements on large distances. Paper presented at the Experimental Vibration Analysis for Civil Engineering Structures Conference, Oura Preta, Brazil.Google Scholar
Papathanassoglou, E. D. E. & Mpouzika, M. D. A. (2012). Interpersonal touch: physiological effects in critical care. Biological Research for Nursing, 14: 431443.Google Scholar
Parati, G., Pomidossi, G., Casadei, R., & Mancia, G. (1985). Lack of alerting reactions to intermittent cuff inflations during noninvasive blood pressure monitoring. Hypertension, 7: 597601.Google Scholar
Paukkunen, M., Parkkila, P., Kettunen, R., & Sepponen, R. (2015). Unified frame of reference improves inter-subject variability of seismocardiograms. Biomedical Engineering Online, 14: 16.Google Scholar
Pauling, J. D., Shipley, J. A., Raper, S., Watson, M. L., Ward, S. G., Harris, N. D., & McHugh, N. J. (2012). Comparison of infrared thermography and laser speckle contrast imaging for the dynamic assessment of digital microvascular function. Microvascular Research, 83: 162167.Google Scholar
Paulson, C. N., Chang, J. T., Romero, C. E., Watson, J., Pearce, F. J., & Levin, N. (2005). Ultra-wideband radar methods and techniques of medical sensing and imaging. Paper presented at the Optics East Conference.Google Scholar
Peck, E. M., Afergan, D., Yuksel, B. F., Lalooses, F., & Jacob, R. J. K. (2014). Using fNIRS to measure mental workload in the real world. In Fairclough, S. H. & Gilleade, K. (eds.), Advances in Physiological Computing (pp. 117139). Heidelberg: Springer.Google Scholar
Pereira, T., Cabeleira, M., Matos, P., Borges, E., Almeida, V., Pereira, H. C., … & Correia, C. (2013). Non-contact pulse wave velocity assessment using optical methods. In Fred, A., Philipe, J., & Gamboa, H. (eds.), Biomedical Engineering Systems and Technologies (pp. 246257). Berlin: Springer.Google Scholar
Perloff, J. K. (2000). Physical Examination of the Heart and Circulation, 3rd edn. Philadelphia, PA: W. B. Saunders.Google Scholar
Picard, R. (1997). Affective Computing. Cambridge, MA: MIT Press.Google Scholar
Picton, T. W. & Hillyard, S. A. (1972). Cephalic skin potentials in electroencephalography. Electroencephalography & Clinical Neurophysiology, 33: 419424.Google Scholar
Pietrangelo, S. J. (2013). An Electronically Steered, Wearable Transcranial Doppler Ultrasound System. Cambridge, MA: Massachusetts Institute of Technology.Google Scholar
Pinheiro, E, Postolache, O., & Girao, P. (2010). Theory and developments in an unobtrusive cardiovascular system representation: ballistocardiography. Open Biomedical Engineering Journal, 4: 201216.Google Scholar
Pinheiro, E., Postolache, O., & Girao, P. (2013). Contactless impedance cardiography using embedded sensors. Measurement Science Review, 13: 157164.Google Scholar
Pino, C. & Kavasidis, I. (2012). Improving mobile device interaction by eye tracking analysis. Paper presented at the Federated Conference on Computer Science and Information Systems (FedCSIS).Google Scholar
Poh, M.-Z., McDuff, D. J., & Picard, R. W. (2010a). Non-contact, automated cardiac pulse measurements using video imaging and blind source separation. Optics Express, 18: 1076210774.Google Scholar
Poh, M.-Z., McDuff, D. J., & Picard, R. W. (2011). Advancements in noncontact, multiparameter physiological measurements using a webcam. IEEE Transactions on Biomedical Engineering, 58: 711.Google Scholar
Poh, M.-Z., Swenson, N. C., & Picard, R. W. (2010b). A wearable sensor for unobtrusive, long-term assessment of electrodermal activity. IEEE Transactions on Biomedical Engineering, 57: 12431252.Google Scholar
Poon, C. C. Y., Zhang, Y.-T., Wong, G., & Poon, W. S. (2008). The beat-to-beat relationship between pulse transit time and systolic blood pressure. Paper presented at the International Conference on Information Technology and Applications in Biomedicine.Google Scholar
Popovic, Z., Momenroodaki, P., & Scheeler, R. (2014). Toward wearable wireless thermometers for internal body temperature measurements. IEEE Communications Magazine, 52: 118125.Google Scholar
Protogerou, A. D., Argyris, A., Nasothimiou, E., Vrachatis, D., Papaioannou, T. G., Tzamouranis, D., … & Stergiou, G. S. (2012). Feasibility and reproducibility of noninvasive 24-h ambulatory aortic blood pressure monitoring with a brachial cuff-based oscillometric device. American Journal of Hypertension, 25: 876882.Google Scholar
Protogerou, A. D., Argyris, A. A., Papaioannou, T. G., Kollias, G. E., Konstantonis, G. D., Nasothimiou, E., … & Sfikakis, P. P. (2014). Left-ventricular hypertrophy is associated better with 24-h aortic pressure than 24-h brachial pressure in hypertensive patients: the SAFAR study. Journal of Hypertension, 32: 18051814.Google Scholar
Ramachandran, G., Swarnamani, S., & Singh, M. (1991). Reconstruction of out-of-plane cardiac displacement patterns as observed on the chest wall during various phases of ECG by capacitance transducer. IEEE Transactions on Biomedical Engineering, 38: 383385.Google Scholar
Rasooly, A. & Herold, K. E. (2015). Mobile Health Technologies: Methods and Protocols. New York: Springer.Google Scholar
Rebolledo-Mendez, G., Reyes, A., Paszkowicz, S., Domingo, M. C., & Skrypchuk, L. (2014). Developing a body sensor network to detect emotions during driving. IEEE Transactions on Intelligent Transportation Systems, 15: 18501854.Google Scholar
Redon, J. & Lurbe, E. (2014). Ambulatory blood pressure monitoring is ready to replace clinic blood pressure in the diagnosis of hypertension: con side of the argument. Hypertension, 64: 11691174.Google Scholar
Rickles, W. H. J. & Day, J. L. (1968). Electrodermal activity in non-palmer skin sites. Psychophysiology, 4: 421435.Google Scholar
Ring, E. F. J. (2006). The historical development of thermometry and thermal imaging in medicine. Journal of Medical Engineering & Technology, 30: 192198.Google Scholar
Roddie, I. C. (1977). Human responses to emotional stress. Irish Journal of Medical Science, 146: 395417.Google Scholar
Rogers, J. A. (2015). Electronics for the human body. Journal of the American Medical Association, 313: 561562.Google Scholar
Rohrbaugh, J. W., Sirevaag, E. J., & Richter, E. J. (2013). Laser Doppler vibrometry measurement of the mechanical myogram. Review of Scientific Instruments, 84: 121706.Google Scholar
Rose, D. P., Ratterman, M., Griffin, D. K., Hou, L., Kelley-Loughnane, N., Naik, R. R., … & Heikenfeld, J. (2014). Adhesive RFID sensor patch for monitoring of sweat electrolytes. IEEE Transactions on Biomedical Engineering, 62: 14571465.Google Scholar
Sadeh, A., Hauri, P. J., Kripke, D. F., & Lavie, P. (1995). The role of actigraphy in the evaluation of sleep disorders. Sleep, 18: 288302.Google Scholar
Safrai, E., Ishai, P. B., Caduff, A., Puzenko, A., Polsman, A., Agranat, A. J., & Feldman, Y. (2012). The remote sensing of mental stress from the electromagnetic reflection coefficient of human skin in the sub-THz range. Bioelectromagnetics, 33: 375382.Google Scholar
Saito, A., Ianov, A., & Sankai, Y. (2009). Measurement of brain activity using optical and electrical method. Paper presented at the IEEE International Conference on Robotics and Biomimetics (ROBIO).Google Scholar
Sarbin, T. R. (1944). The logic of prediction in psychology. Psychological Review, 51: 210228.Google Scholar
Sato, K., Kang, W., Saga, K., & Sato, K. T. (1989). Biology of sweat glands and their disorders: I. Normal sweat gland function. Journal of the American Academy of Dermatology, 20: 537563.Google Scholar
Sato, K. & Sato, F. (1983). Individual variations in structure and function of human eccrine sweat gland. American Journal of Physiology, 14: R203R208.Google Scholar
Scalise, L. (2012). Non contact heart monitoring. In Millis, R. (ed.), Advances in Electrocardiograms: Methods and Analysis (pp. 81106). Rijeka, Croatia: InTech.Google Scholar
Scalise, L., Casaccia, S., Marchionni, P., Ercoli, I., & Tomasini, E. P. (2013). Laser Doppler myography (LDMi): a novel non-contact measurement method for the muscle activity. Laser Therapy, 22: 261268.Google Scholar
Scalise, L., Ercoli, I., & Marchionni, P. (2010). Optical method for measurement of respiration rate. Paper presented at the IEEE International Workshop on Medical Measurements and Applications, Ottawa, April 30–May 1.Google Scholar
Schafer, A. & Vagedes, J. (2013). How accurate is pulse rate variability as an estimate of heart rate variabilty? A review on studies comparing photoplethysmographic technology with an electrocardiogram. International Journal of Cardiology, 166: 1529.Google Scholar
Schlotz, W. (2012). Ambulatory psychoneuroendocrinology: assessing salivary cortisol and other hormones in daily life. In Mehl, M. R. & Conner, T. S. (eds.), Handbook of Research Methods for Studying Daily Life (pp. 193209). New York: Guilford Press.Google Scholar
Scully, C., Lee, J., Meyer, J., Gorbach, A. M., Granquist-Fraser, D., Mendelson, Y., & Chon, K. H. (2012). Physiological parameter monitoring from optical recordings with a mobile phone. IEEE Transactions on Biomedical Engineering, 59: 303306.Google Scholar
Searle, A. & Kirkup, L. (2000). A direct comparison of wet, dry and insulating bioelectric recording electrodes. Physiological Measurement, 21: 271283.Google Scholar
Secco, E. L., Curone, D., Tognetti, A., Bonfiglio, A., & Magenes, G. (2012). Validation of smart garments for physiological and activity-related monitoring of humans in harsh environment. American Journal of Biomedical Engineering, 2: 189196.Google Scholar
Sewell, W. & Komogortsev, O. (2010). Real-time eye gaze tracking with an unmodified commodity webcam employing a neural network. Paper presented at the CHI’10 Extended Abstracts on Human Factors in Computing Systems.Google Scholar
Shafiq, G. & Veluvolu, K. C. (2014). Surface chest motion decomposition for cardiovascular monitoring. Scientific Reports, 4: 5093.Google Scholar
Shao, D., Yang, Y., Liu, C., Tsow, F., Yu, H., & Tao, N. (2014). Non-contact monitoring breathing pattern, exhalation flow rate and pulse transit time. IEEE Transactions on Biomedical Engineering, 61: 27602767.Google Scholar
Sherwood, A. & Turner, J. R. (1993). Postural stability of hemodynamic responses during mental challenge. Psychophysiology, 30: 237244.Google Scholar
Shin, J. H., Lee, K. M., & Park, K. S. (2009). Non-constrained monitoring of systolic blood pressure on a weighing scale. Physiological Measurement, 30: 679693.Google Scholar
Silverman, I. (1974). The experimenter: a (still) neglected stimulus object. Canadian Psychologist, 15: 258270.Google Scholar
Singh, M. & Ramachandran, G. (1991). Reconstruction of sequential cardiac in-plane displacement patterns on the chest wall by laser speckle interferometry. IEEE Transactions on Biomedical Engineering, 38: 483489.Google Scholar
Sirevaag, E. J., Casaccia, S., Richter, E. J., O’Sullivan, J. A., Scalise, L., & Rohrbaugh, J. W. (2016). Cardio-respiratory interactions: non-contact assessment using laser Doppler vibrometry. Psychophysiology, 53: 847867.Google Scholar
Sloan, R. P., Shapiro, P. A., Bagiella, E., Fishkin, P. E., Gorman, J. M., & Myers, M. M. (1995). Consistency of heart rate and sympathovagal reactivity across different autonomic contexts. Psychophysiology, 32: 452459.Google Scholar
Smith, N. T. (1974). Ballistocardiography. In Weissler, A. M. (ed.), Noninvasive Cardiology (pp. 39148). New York: Grune & Stratton.Google Scholar
Sola, J., Proenca, M., Ferrario, D., Porchet, J.-A., Falhi, A., Grossenbacher, O., … & Sartori, C. (2013). Noninvasive and nonocclusive blood pressure estimation via a chest sensor. IEEE Transactions on Biomedical Engineering, 60: 35053513.Google Scholar
Sparks, S. A., Chandler, P., Bailey, T. G., Marchant, D. C., & Orme, D. (2013). The energy demands of portable gas analysis system carriage during walking and running. Ergonomics, 56: 19011907.Google Scholar
Squara, P. (2008). Bioreactance: a new method for non-invasive cardiac output monitoring. In Vincent, J.-L. (ed.), Yearbook of Intensive Care and Emergency Medicine (pp. 619630). Berlin: Springer.Google Scholar
Steffen, M., Aleksandrowicz, A., & Leonhardt, S. (2007). Mobile noncontact monitoring of heart and lung activity. IEEE Transactions on Biomedical Circuits and Systems, 1: 250257.Google Scholar
Stokes, M. & Blythe, M. (2001). Muscle Sounds in Physiology, Sports Science and Clinical Investigation. Oxford: Medintel.Google Scholar
Stopczynski, A., Stahlhut, C., Larsen, J. E., Petersen, M. K., & Hansen, L. K. (2014). The smartphone brain scanner: a portable real-time neuroimaging system. PloS One, 9: e86733.Google Scholar
Sun, Y., Papin, C., Azorin-Peris, V., Kalawsky, R., Greenwald, S., & Hu, S. (2012). Use of ambient light in remote photoplethysmographic systems: comparison between a high-performance camera and a low-cost webcam. Journal of Biomedical Optics, 17: 037005103700510.Google Scholar
Sutarman, & Thomson, M. L. (1952). A new technique for enumerating active sweat glands in man. Journal of Physiology, 117: 51P52P.Google Scholar
Suzuki, S., Matsui, T., Sugawara, K., Asao, T., & Kotani, K. (2011). An approach to remote monitoring of heart rate variability (HRV) using microwave radar during a calculation task. Journal of Physiological Anthropology, 30: 241249.Google Scholar
Takano, C. & Ohta, Y. (2007). Heart rate measurement based on a time-lapse image. Medical Engineering & Physics, 29: 853857.Google Scholar
Tamura, T., Maeda, Y., Sekine, M., & Yoshida, M. (2014). Wearable photoplethysmographic sensors: past and present. Electronics, 3: 282302.Google Scholar
Tanaka, S., Gao, S., Nogawa, M., & Yamakoshi, K.-I. (2005). Noninvasive measurement of instantaneous, radial artery blood pressure. IEEE Engineering in Medicine and Biology Magazine, 24: 3237.Google Scholar
Tapia, E. M., Intille, S. S., Haskell, W., Larson, K., Wright, J., King, A., & Friedman, R. (2007). Real-time recognition of physical activities and their intensities using wireless accelerometers and a heart rate monitor. Paper presented at the 11th IEEE International Symposium on Wearable Computers.Google Scholar
Tartz, R., Vartak, A., King, J., & Fowles, D. C. (2015). Effects of grip force on skin conductance measured from a handheld device. Psychophysiology, 52: 819.Google Scholar
Tassinary, L. G., Cacioppo, J. T., & Vanman, E. J. (2007). The skeletomotor system: Surface electromyography. In Cacioppo, J. T., Tassinary, L. G., & Berntson, G. (eds.), Handbook of Psychophysiology, 3rd edn. (pp. 267299). Cambridge University Press.Google Scholar
Tavakolian, K., Dumont, G., & Blaber, A. (2012). Analysis of seismocardiogram capability for trending stroke volume changes: a lower body negative pressure study. Paper presented at the Computing in Cardiology (CinC) Conference.Google Scholar
Tavakolian, K., Ngai, B., Blaber, A. P., & Kaminska, B. (2011). Infrasonic cardiac signals: complementary windows to cardiovascular dynamics. Paper presented at the Engineering in Medicine and Biology Society, Annual International Conference of the IEEE.Google Scholar
Tavel, M. E. (1972). Clinical Phonocardiography and External Pulse Recording, 2nd edn. Chicago: Year Book Medical Publishers.Google Scholar
Teichmann, D., De Matteis, D., Walter, M., & Leonhardt, S. (2014). A bendable and wearable cardiorespiratory monitoring device fusing two noncontact sensor principles. Paper presented at the 11th International Conference on Wearable and Implantable Body Sensor Networks (BSN).Google Scholar
Tharion, W. J., Buller, M. J., Potter, A. W., Karis, A. J., Goetz, V., & Hoyt, R. W. (2013). Acceptability and usability of an ambulatory health monitoring system for use by military personnel. IIE Transactions on Occupational Ergonomics and Human Factors, 1: 203214.Google Scholar
Thiel, F., Kreiseler, D., & Seifert, F. (2009). Non-contact detection of myocardium’s mechanical activity by ultrawideband RF-radar and interpretation applying electrocardiography. Review of Scientific Instruments, 80: 114302.Google Scholar
Thomas, P. E. & Korr, I. M. (1957). Relationship between sweat gland activity and electrical resistance of the skin. Journal of Applied Physiology, 10: 505510.Google Scholar
Tripathi, S. R., Miyata, E., Ishai, P. B., & Kawase, K. (2015). Morphology of human sweat ducts observed by optical coherence tomography and their frequency of resonance in the terahertz frequency region. Scientific Reports, 5: 9071.Google Scholar
Trobec, R., Rashkovska, A., & Avebelj, V. (2012). Two proximal skin electrodes: A respiration rate sensor. Sensors, 12: 1381313828.Google Scholar
Trull, T. J. & Ebner-Priemer, U. (2013). Ambulatory assessment. Annual Review of Clinical Psychology, 9: 151176.Google Scholar
Turner, J. R., Sherwood, A., & Light, K. C. (1991). Generalization of cardiovascular response: supportive evidence for the reactivity hypothesis. International Journal of Psychophysiology, 11: 207212.Google Scholar
Ulbrich, M., Muhlsteff, J., Sipila, A., Kamppi, M., Koskela, A., Myry, M., … & Walter, M. (2014). The IMPACT shirt: textile integrated and portable impedance cardiography. Physiological Measurement, 35: 11811196.Google Scholar
Vainer, B. G. (2005). FPA-based infrared thermography as applied to the study of cutaneous perspiration and stimulated vascular response in humans. Physics in Medicine and Biology, 50: R63R94.Google Scholar
Van Lange, P. A. M., Finkenauer, C., Popma, A., & van Vugt, M. (2011). Electrodes as social glue: measuring heart rate promotes giving in the trust game. International Journal of Psychophysiology, 80: 246250.Google Scholar
Van Lien, R., Neijts, M., Willemsen, G., & De Gues, E. J. C. (2015). Ambulatory measurement of the ECG T-wave amplitude. Psychophysiology, 52: 225237.Google Scholar
Vassend, O. & Knardahl, S. (2005). Personality, affective response, and facial blood flow during brief cognitive tasks. International Journal of Psychophysiology, 55: 265278.Google Scholar
Verkruysse, W., Svaasand, L. O., & Nelson, J. S. (2008). Remote plethysmographic imaging using ambient light. Optics Express, 16: 2143421445.Google Scholar
Vlachopoulos, C., Alexopoulos, N., & Stefanadis, C. (2006). Lifestyle modification and arterial stiffness and wave reflections: a more natural way to prolong arterial health. Artery Research, 1: S15S22.Google Scholar
Vlachopoulos, C., Aznaouridis, K., & Stefanadis, C. (2010). Prediction of cardiovascular events and all-cause mortality with arterial sltiffness: a systematic review and meta-analysis. Journal of the American College of Cardiology, 55: 13181327.Google Scholar
Vlachopoulos, C., Xaplanteris, P., Alexopoulos, N., Aznaouridis, K., Vasiliadou, C., Baou, K., … & Stefanadis, C. (2009). Divergent effects of laughter and mental stress on arterial stiffness and central hemodynamics. Psychosomatic Medicine, 71: 446453.Google Scholar
Vural, E., Simske, S., & Schuckers, S. (2013). Verification of individuals from accelerometer measures of cardiac chest movements. Paper presented at the International Conference of the Biometrics Special Interest Group (BIOSIG).Google Scholar
Waldstein, S. R., Neumann, S. A., & Merrill, J. A. (1998). Postural effects on hemodynamic response to interpersonal interaction. Biological Psychology, 48: 5767.Google Scholar
Walsh, J. A. I., Topol, E. J., & Steinhubl, S. R. (2014). Novel wireless devices for cardiac monitoring. Circulation, 130: 573581.Google Scholar
Wang, S. & Zhou, G. (2015). A review on radio based activity recognition. Digital Communications and Networks, 1: 2029.Google Scholar
Wassertheurer, S., Kropf, J., Weber, T., Van der Giet, M., Baulmann, J., Ammer, M., … & Magometschnigg, D. (2010). A new oscillometric method for pulse wave analysis: comparison with a common tonometric method. Journal of Human Hypertension, 24: 498504.Google Scholar
Watanabe, T. & Watanabe, K. (2004). Noncontact method for sleep stage estimation. IEEE Transactions on Biomedical Engineering, 151: 17351748.Google Scholar
Weissler, A. M. (1974). Noninvasive Cardiology. New York: Grune & Stratton.Google Scholar
Wengrowski, E. (2014). A survey on device-free passive localization and gesture recognition via body wave reflections. ACM Transactions on Autonomous and Adaptive Systems, 5: 115.Google Scholar
Westerink, J., van Beek, W., Daemen, E., Janssen, J., de Vries, G.-J., & Ouwerkerk, M. (2014). The vitality bracelet: bringing balance to your life with psychophysiological measurements. In Fairclough, S. H. & Gilleade, K. (eds.), Advances in Physiological Computing (pp. 197209). London: Springer.Google Scholar
Wilcott, R. C. (1960). A comparison of palmar and nonpalmar skin conductance. Journal of Comparative and Physiological Psychology, 53: 3841.Google Scholar
Wilcott, R. C. (1962). Palmar skin sweating vs. palmar skin resistance and skin potential. Journal of Comparative and Physiological Psychology, 55: 327331.Google Scholar
Wilhelm, F. H., Grossman, P., & Muller, M. I. (2012). Bridging the gap between the laboratory and the real world: integrative ambulatory psychophysiology. In Mehl, M. R. & Conner, T. S. (eds.), Handbook of Research Methods for Studying Daily Life (pp. 210234). New York: Guilford Press.Google Scholar
Wilhelm, F. H., Roth, W. T., & Sackner, M. A. (2003). The LifeShirt: an advanced system for ambulatory measurement of respiratory and cardiac function. Behavior Modification, 27: 671691.Google Scholar
Wilkin, J. K. (1988). Why is flushing limited to a mostly facial cutaneous distribution? Journal of the American Academy of Dermatology, 10: 309313.Google Scholar
Wilson, A. D. & Baietto, M. (2011). Advances in electronic-nose technologies developed for biomedical applications. Sensors, 11: 11051176.Google Scholar
Winokur, E. S., He, D. D., & Sodini, C. G. (2012). A wearable vital signs monitor at the ear for continuous heart rate and pulse transit time measurements. Paper presented at the 34th Annual International Conference of the IEEE EMBS, San Diego, CA.Google Scholar
Wollburg, E., Roth, W. T., & Kim, S. (2009). End-tidal transcutaneous measurement of PCO2 during voluntary hypo- and hyperventilation. International Journal of Psychophysiology, 71: 103108.Google Scholar
Wongchoosuk, C., Lutz, M., & Kerdcharoen, T. (2009). Detection and classification of human body odor using an electronic nose. Sensors, 9: 72347249.Google Scholar
Wood, E. & Bulling, A. (2014). Eyetab: model-based gaze estimation on unmodified tablet computers. Paper presented at the Proceedings of the Symposium on Eye Tracking Research and Applications.Google Scholar
Wu, C., Yang, Z., Zhou, Z., Liu, X., Liu, Y., & Cao, J. (2015). Non-invasive detection of moving and stationary human with WiFi. IEEE Journal on Selected Areas in Communications, 33: 1.Google Scholar
Wu, H.-Y., Rubinstein, M., Shih, E., Guttag, J., Durand, F., & Freeman, W. (2012). Eulerian video magnification for revealing subtle changes in the world. ACM Transactions on Graphics, 31.Google Scholar
Xiao, S., Cai, S., & Liu, G. (2000). Studying the significance of cardiac contractility variability. IEEE Engineering in Medicine and Biology Magazine, 19: 102105.Google Scholar
Xiao, Y., Lin, J., Boric-Lubecke, O., & Lubecke, V. M. (2006). A Ka-band low power Doppler radar system for remote detection of cardiopulmonary motion. Paper presented at the Engineering in Medicine and Biology Society, 27th Annual International Conference of the IEEE.Google Scholar
Xu, S., Zhang, Y., Jia, L., Mathewson, K. E., Jang, K.-I., Kim, J., … & Wang, R. (2014). Soft microfluidic assemblies of sensors, circuits, and radios for the skin. Science, 344: 7074.Google Scholar
Yakymenko, I., Tsybulin, O., Sidorik, E., Henshel, D., Kyrylenko, O., & Kyrylenko, S. (2016). Oxidative mechanisms of biological activity of low-intensity radiofrequency radiation. Electromagnetic Biology and Medicine, 35: 186202.Google Scholar
Yeo, W. H., Kim, Y. S., Lee, J., Ameen, A., Shi, L., Li, M., … & Kang, Z. (2013). Multifunctional epidermal electronics printed directly onto the skin. Advanced Materials, 25: 27732778.Google Scholar
Yin, L., Huang, X., Xu, H., Zhang, Y., Lam, J., Cheng, J., & Rogers, J. A. (2014). Materials, designs, and operational characteristics for fully biodegradable primary batteries. Advanced Materials, 26: 38793884.Google Scholar
Zanetti, J. M. & Salerno, D. M. (1991). Seismocardiography: a technique for recording precordial acceleration. Paper presented at the Computer-Based Medical Systems, 4th Annual IEEE Symposium.Google Scholar
Zhang, J., Ser, W., & Goh, D. Y. T. (2011). A novel respiratory rate estimation method for sound-based wearable monitoring systems. Paper presented at the Engineering in Medicine and Biology Society, Annual International Conference of the IEEE.Google Scholar
Zhang, S. L., Meyers, C. L., Subramanyan, K., & Hancewicz, T. (2005). Near infrared imaging for measuring and visualizing skin hydration: a comparison with visual assessment and electrical methods. Journal of Biomedical Optics, 10: 031107.Google Scholar
Zhao, F., Li, M., Qian, Y., & Tsien, J. Z. (2013). Remote measurements of heart and respiration rates for telemedicine. PloS One, 8: e71384.Google Scholar
Zheng, J., Hu, S., Azorin-Peris, V., Echiadis, A., Chouliaras, V., & Summers, R. (2008). Remote simultaneous dual wavelength imaging photoplethysmography: a further step towards 3-D mapping of skin blood microcirculation. Paper presented at the Multimodal Biomedical Imaging III.Google Scholar
Zheng, Y.-L., Yan, B. P., Zhang, Y.-T., & Poon, C. C. Y. (2014). An armband wearable device for overnight and cuff-less blood pressure measurement. IEEE Transactions on Biomedical Engineering, 61: 21792186.Google Scholar
Zhou, D., Luo, J., Silenzio, V., Zhou, Y., Hu, J., Currier, G., & Kautz, H. (2015). Tackling mental health by integrating unobtrusive multimodal sensing. Paper presented at the 29th AAAI Conference on Artificial Intelligence.Google Scholar
Zito, D., Pepe, D., Mincica, M., Zito, F., De Rossi, D., Scilingo, E. P., & Tognetti, A. (2008). Wearable system-on-a-chip UWB radar for contact-less cardiopulmonary monitoring: present status. Paper presented at the Engineering in Medicine and Biology Society, 30th Annual International Conference of the IEEE.Google Scholar

References

Andreassen, O. A., Thompson, W. K., Schork, A. J., Ripke, S., Mattingsdal, M., Kelsoe, J. R., … & Dale, A. M. (2013). Improved detection of common variants associated with schizophrenia and bipolar disorder using pleiotropy-informed conditional false discovery rate. PLoS Genetics, 9: e1003455.Google Scholar
Anokhin, A., Steinlein, O., Fischer, C., Mao, Y., Vogt, P., Schalt, E., & Vogel, F. (1992). A genetic study of the human low-voltage electroencephalogram. Human Genetics, 90: 99112.Google Scholar
Anokhin, A. P. (2014). Genetic psychophysiology: Advances, problems, and future directions. International Journal of Psychophysiology, 93: 173197.Google Scholar
Begleiter, H., Porjesz, B., Reich, T., Edenberg, H. J., Goate, A., Blangero, J., … & Bloom, F. E. (1998). Quantitative trait loci analysis of human event-related brain potentials: P3 voltage. Electroencephalography & Clinical Neurophysiology/Evoked Potentials Section, 108: 244250.Google Scholar
Blokland, G. A., de Zubicaray, G. I., McMahon, K. L., & Wright, M. J. (2012). Genetic and environmental influences on neuroimaging phenotypes: a meta-analytical perspective on twin imaging studies. Twin Research and Human Genetics, 15: 351371.Google Scholar
Boker, S., Neale, M., Maes, H., Wilde, M., Spiegel, M., Brick, T., … & Fox, J. (2011). OpenMx: an open source extended structural equation modeling framework. Psychometrika, 76: 306317.Google Scholar
Bollen, K. A. (1989). Structural Equations with Latent Variables. New York: John Wiley.Google Scholar
Boomsma, D., Busjahn, A., & Peltonen, L. (2002). Classical twin studies and beyond. Nature Reviews Genetics, 3: 872882.Google Scholar
Boomsma, D. I., De Geus, E. J., Van Baal, G. C., & Koopmans, J. R. (1999). A religious upbringing reduces the influence of genetic factors on disinhibition: evidence for interaction between genotype and environment on personality. Twin Research, 2: 115125.Google Scholar
Boomsma, D. I. & Gabrielli, W. F. (1985). Behavioral genetic approaches to psychophysiological data. Psychophysiology, 22: 249260.Google Scholar
Boomsma, D. I. & Molenaar, P. C. (1986). Using LISREL to analyze genetic and environmental covariance structure. Behavior Genetics, 16: 237250.Google Scholar
Boomsma, D. I. & Molenaar, P. C. (1987). The genetic analysis of repeated measures: I. Simplex models. Behavior Genetics, 17: 111123.Google Scholar
Bouchard, T. J. & McGue, M. (2003). Genetic and environmental influences on human psychological differences. Journal of Neurobiology, 54: 445.Google Scholar
Browne, M. W. (1973). Generalized least squares estimators in the analysis of covariance structures. ETS Research Bulletin Series, 1: 136.Google Scholar
Bulik-Sullivan, B., Finucane, H. K., Anttila, V., Gusev, A., Day, F. R., Perry, J. R., … & Neale, B. M. (2015a). An atlas of genetic correlations across human diseases and traits. Nature Genetics, 47: 12361241.Google Scholar
Bulik-Sullivan, B., Loh, P. R., Finucane, H., Ripke, S., Yang, J., Patterson, N., … & Neale, B. M. (2015b). LD score regression distinguishes confounding from polygenicity in genome-wide association studies. Nature Genetics, 47: 291295.Google Scholar
Carey, G. (1986). Sibling imitation and contrast effects. Behavior Genetics, 16: 319341.Google Scholar
Cederlof, R., Friberg, L., & Lundman, T. (1977). The interactions of smoking, environment and heredity and their implications for disease etiology: a report of epidemiological studies on the Swedish twin registries. Acta Medica Scandinavica. Supplementum, 612: 1128.Google Scholar
Cloninger, C. R., Rice, J., & Reich, T. (1979). Multifactorial inheritance with cultural transmission and assortative mating: II. A general model of combined polygenic and cultural inheritance. American Journal of Human Genetics, 31: 176198.Google Scholar
De Bakker, P. I., Ferreira, M. A., Jia, X., Neale, B. M., Raychaudhuri, S., & Voight, B. F. (2008). Practical aspects of imputation-driven meta-analysis of genome-wide association studies. Human Molecular Genetics, 17: R122R128.Google Scholar
De Moor, M. H., Boomsma, D. I., Stubbe, J. H., Willemsen, G., & De Geus, E. J. (2008). Testing causality in the association between regular exercise and symptoms of anxiety and depression. Archives of General Psychiatry, 65: 897905.Google Scholar
de Zeeuw, E. L., van Beijsterveldt, C. E., Glasner, T. J., Bartels, M., Ehli, E. A., Davies, G. E., … & Boomsma, D. I. (2014). Polygenic scores associated with educational attainment in adults predict educational achievement and ADHD symptoms in children. American Journal of Medical Genetics Part B: Neuropsychiatric Genetics, 165: 510520.Google Scholar
den Braber, A., Bohlken, M. M., Brouwer, R. M., van’t Ent, D., Kanai, R., Kahn, R. S., … & Boomsma, D. I. (2013). Heritability of subcortical brain measures: a perspective for future genome-wide association studies. NeuroImage, 83: 98102.Google Scholar
den Hoed, M., Eijgelsheim, M., Esko, T., Brundel, B. J., Peal, D. S., Evans, D. M., … & Loos, R. J. (2013). Identification of heart rate-associated loci and their effects on cardiac conduction and rhythm disorders. Nature Genetics, 45: 621631.Google Scholar
Dolan, C. V., de Kort, J. M., Van Beijsterveldt, T. C., Bartels, M., & Boomsma, D. I. (2014). GE covariance through phenotype to environment transmission: an assessment in longitudinal twin data and application to childhood anxiety. Behavior Genetics, 44: 240253.Google Scholar
Dudbridge, F. (2013). Power and predictive accuracy of polygenic risk scores. PLoS Genetics, 9: e1003348.Google Scholar
Eaves, L. J. (1976). A model for sibling effects in man. Heredity, 36: 205214.Google Scholar
Eaves, L. J. (1987). Including the environment in models for genetic segregation. Journal of Psychiatric Research, 21: 639647.Google Scholar
Ehli, E. A., Abdellaoui, A., Hu, Y., Hottenga, J. J., Kattenberg, M., van Beijsterveldt, T., … & Davies, G. E. (2012). De novo and inherited CNVs in MZ twin pairs selected for discordance and concordance on attention problems. European Journal of Human Genetics, 20: 10371043.Google Scholar
Euesden, J., Lewis, C. M., & O’Reilly, P. F. (2014). PRSice: Polygenic Risk Score software. Bioinformatics, 31: 14661468.Google Scholar
Falconer, D. S. & Mackay, T. F. (1996). Introduction to Quantitative Genetics. Harlow: Longman.Google Scholar
Finucane, H. K., Bulik-Sullivan, B., Gusev, A., Trynka, G., Reshef, Y., Loh, P. R., … & Price, A. L. (2015). Partitioning heritability by functional category using GWAS summary statistics. Nature Genetics, 47: 12281235.Google Scholar
Fisher, R. A. (1918). The correlation of relatives on the assumption of Mendelian inheritance. Philosophical Transactions of the Royal Society of Edinburgh, 52: 399433.Google Scholar
Franić, S., Dolan, C. V., Broxholme, J., Hu, H., Zemojtel, T., Davies, G. E., … & Boomsma, D. I. (2015). Mendelian and polygenic inheritance of intelligence: a common set of causal genes? Using next-generation sequencing to examine the effects of 168 intellectual disability genes on normal-range intelligence. Intelligence, 49: 1022.Google Scholar
Freund, J., Brandmaier, A. M., Lewejohann, L., Kirste, I., Kritzler, M., Krüger, A., … & Kempermann, G. (2013). Emergence of individuality in genetically identical mice. Science, 340: 756759.Google Scholar
Fulker, D. W., Baker, L. A., & Bock, R. D. (1983). Estimating components of covariance using LISREL. Data Analyst, 1: 58.Google Scholar
Genomes Project Consortium (2012). An integrated map of genetic variation from 1,092 human genomes. Nature, 491: 5665.Google Scholar
Groen-Blokhuis, M. M., Middeldorp, C. M., van Beijsterveldt, C. E., & Boomsma, D. I. (2011). Evidence for a causal association of low birth weight and attention problems. Journal of the American Academy of Child & Adolescent Psychiatry, 50: 12471254.Google Scholar
Gusev, A., Lee, S. H., Neale, B. M., Trynka, G., Vilhjalmsson, B. J., Finucane, H., … & Price, A. L. (2014). Regulatory variants explain much more heritability than coding variants across 11 common diseases. bioRxiv, 004309. doi: http://dx.doi.org/10.1101/004309Google Scholar
Hamer, D. H. & Sirota, L. (2000). Beware the chopsticks gene. Molecular Psychiatry, 5: 1113.Google Scholar
Haseman, J. K. & Elston, R. C. (1972). The investigation of linkage between a quantitative trait and a marker locus. Behavior Genetics, 2: 319.Google Scholar
Hewitt, J. K., Eaves, L. J., Neale, M. C., & Meyer, J. M. (1988). Resolving causes of developmental continuity or “tracking”: I. Longitudinal twin studies during growth. Behavior Genetics, 18: 133151.Google Scholar
Hibar, D. P., Stein, J. L., Renteria, M. E., Arias-Vasquez, A., Desrivières, S., Jahanshad, N., … & Medland, S. E. (2015). Common genetic variants influence human subcortical brain structures. Nature, 520: 224229.Google Scholar
Hodgkinson, C. A., Enoch, M. A., Srivastava, V., Cummins-Oman, J. S., Ferrier, C., Iarikova, P., … & Goldman, D. (2010). Genome-wide association identifies candidate genes that influence the human electroencephalogram. Proceedings of the National Academy of Sciences of the USA, 107: 86958700.Google Scholar
Hoggart, C. J., Clark, T. G., De Iorio, M., Whittaker, J. C., & Balding, D. J. (2008). Genome-wide significance for dense SNP and resequencing data. Genetic Epidemiology, 32: 179185.Google Scholar
Hottenga, J. J., Whitfield, J. B., Posthuma, D., Willemsen, G., De Geus, E. J., Martin, N. G., & Boomsma, D. I. (2007). Genome-wide scan for blood pressure in Australian and Dutch subjects suggests linkage at 5P, 14Q, and 17P. Hypertension, 49: 832838.Google Scholar
Howie, B., Fuchsberger, C., Stephens, M., Marchini, J., & Abecasis, G. R. (2012). Fast and accurate genotype imputation in genome-wide association studies through pre-phasing. Nature Genetics, 44: 955959.Google Scholar
Iacono, W. G. (2014). Genome-wide scans of genetic variants for psychophysiological endophenotypes: introduction to this special issue of Psychophysiology. Psychophysiology, 51: 12011202.Google Scholar
International HapMap Consortium (2005). A haplotype map of the human genome. Nature, 437: 12991320.Google Scholar
Jinks, J. L. & Fulker, D. W. (1970). Comparison of the biometrical genetical, MAVA, and classical approaches to the analysis of the human behavior. Psychological Bulletin, 73: 311.Google Scholar
Kan, K. J., Dolan, C. V., Nivard, M. G., Middeldorp, C. M., van Beijsterveldt, C. E., Willemsen, G., & Boomsma, D. I. (2013). Genetic and environmental stability in attention problems across the lifespan: evidence from the Netherlands twin register. Journal of the American Academy of Child & Adolescent Psychiatry, 52: 1225.Google Scholar
Keller, M. C. & Coventry, W. L. (2005). Quantifying and addressing parameter indeterminacy in the classical twin design. Twin Research and Human Genetics, 8: 201213.Google Scholar
Keller, M. C., Medland, S. E., Duncan, L. E., Hatemi, P. K., Neale, M. C., Maes, H. H., & Eaves, L. J. (2009). Modeling extended twin family data: I. Description of the Cascade model. Twin Research and Human Genetics, 12: 818.Google Scholar
Kendler, K. S. & Eaves, L. J. (1986). Models for the joint effect of genotype and environment on liability to psychiatric illness. American Journal of Psychiatry, 143: 279289.Google Scholar
Kendler, K. S., Neale, M. C., MacLean, C. J., Heath, A. C., Eaves, L. J., & Kessler, R. C. (1993). Smoking and major depression: a causal analysis. Archives of General Psychiatry, 50: 3643.Google Scholar
Kochunov, P., Glahn, D., Winkler, A., Duggirala, R., Olvera, R. L., Cole, S., … & Blangero, J. (2009). Analysis of genetic variability and whole genome linkage of whole-brain, subcortical, and ependymal hyperintense white matter volume. Stroke, 40: 36853690.Google Scholar
Kruglyak, L. & Lander, E. S. (1995). Complete multipoint sib-pair analysis of qualitative and quantitative traits. American Journal of Human Genetics, 57: 439454.Google Scholar
Lee, S. H., Yang, J., Goddard, M. E., Visscher, P. M., & Wray, N. R. (2012). Estimation of pleiotropy between complex diseases using single-nucleotide polymorphism-derived genomic relationships and restricted maximum likelihood. Bioinformatics, 28: 25402542.Google Scholar
Lippert, C., Listgarten, J., Liu, Y., Kadie, C. M., Davidson, R. I., & Heckerman, D. (2011). FaST linear mixed models for genome-wide association studies. Nature Methods, 8: 833835.Google Scholar
Lubke, G. H., Hottenga, J. J., Walters, R., Laurin, C., De Geus, E. J., Willemsen, G., … & Boomsma, D. I. (2012). Estimating the genetic variance of major depressive disorder due to all single nucleotide polymorphisms. Biological Psychiatry, 72: 707709.Google Scholar
Lynch, M. & Walsh, B. (1998). Genetics and Analysis of Quantitative Traits. Sunderland, MA: Sinauer Associates.Google Scholar
Maes, H. H., Neale, M. C., Medland, S. E., Keller, M. C., Martin, N. G., Heath, A. C., & Eaves, L. J. (2009). Flexible Mx specification of various extended twin kinship designs. Twin Research and Human Genetics, 12: 2634.Google Scholar
Manolio, T. A., Rodriguez, L. L., Brooks, L., Abecasis, G., Ballinger, D., Daly, M., … & Collins, F. S. (2007). New models of collaboration in genome-wide association studies: the Genetic Association Information Network. Nature Genetics, 39: 10451051.Google Scholar
Marchini, J., Howie, B., Myers, S., McVean, G., & Donnelly, P. (2007). A new multipoint method for genome-wide association studies by imputation of genotypes. Nature Genetics, 39: 906913.Google Scholar
Martin, N. G., Boomsma, D. I., & Machin, G. (1997). A twin-pronged attack on complex traits. Nature Genetics, 17: 387392.Google Scholar
Martin, N. G. & Eaves, L. J. (1977). The genetical analysis of covariance structure. Heredity, 38: 7995.Google Scholar
Mather, K. & Jinks, J. L. (1977). Introduction to Biometrical Genetics. Cambridge University Press.Google Scholar
McArdle, J. J. (2006). Latent curve analyses of longitudinal twin data using a mixed-effects biometric approach. Twin Research and Human Genetics, 9: 343359.Google Scholar
McGue, M. & Bouchard, T. Jr. J. (1998). Genetic and environmental influences on human behavioral differences. Annual Review of Neuroscience, 21: 124.Google Scholar
Medland, S. E., Neale, M. C., Eaves, L. J., & Neale, B. M. (2009). A note on the parameterization of Purcell’s G×E model for ordinal and binary data. Behavior Genetics, 39: 220229.Google Scholar
Minicâ, C. C., Dolan, C. V., Kampert, M. M., Boomsma, D. I., & Vink, J. M. (2014). Sandwich corrected standard errors in family-based genome-wide association studies. European Journal of Human Genetics, 23: 388394.Google Scholar
Molenaar, D. & Dolan, C. V. (2014). Testing systematic genotype by environment interactions using item level data. Behavior Genetics, 44: 212231.Google Scholar
Molenaar, D., van der Sluis, S., Boomsma, D. I., Haworth, C. M., Hewitt, J. K., Martin, N. G., … & Dolan, C. V. (2013). Genotype by environment interactions in cognitive ability: a survey of 14 studies from four countries covering four age groups. Behavior Genetics, 43: 208219.Google Scholar
Molenaar, P. C., Boomsma, D. I., & Dolan, C. V. (1993). A third source of developmental differences. Behavior Genetics, 23: 519524.Google Scholar
Nyholt, D. R. (2014). SECA: SNP effect concordance analysis using genome-wide association summary results. Bioinformatics, 30: 20862088.Google Scholar
Pe’er, I., Yelensky, R., Altshuler, D., & Daly, M. J. (2008). Estimation of the multiple testing burden for genomewide association studies of nearly all common variants. Genetic Epidemiology, 32: 381385.Google Scholar
Pennisi, E. (2015). New database links regulatory DNA to its target genes. Science, 348: 618619.Google Scholar
Plomin, R., DeFries, J. C., Knopik, V. S., & Neiderhiser, J. M. (2013). Behavioral Genetics. New York: McMillan Education.Google Scholar
Polderman, T. J., Benyamin, B., de Leeuw, C. A., Sullivan, P. F., van Bochoven, A., Visscher, P. M., & Posthuma, D. (2015). Meta-analysis of the heritability of human traits based on fifty years of twin studies. Nature Genetics, 47: 702709.Google Scholar
Posthuma, D., Beem, A. L., De Geus, E. J., Van Baal, G. C., von Hjelmborg, J. B., Iachine, I., & Boomsma, D. I. (2003). Theory and practice in quantitative genetics. Twin Research, 6: 361376.Google Scholar
Prescott, C. A. (2004). Using the Mplus computer program to estimate models for continuous and categorical data from twins. Behavior Genetics, 34: 1740.Google Scholar
Price, A. L., Patterson, N. J., Plenge, R. M., Weinblatt, M. E., Shadick, N. A., & Reich, D. (2006). Principal components analysis corrects for stratification in genome-wide association studies. Nature Genetics, 38: 904909.Google Scholar
Purcell, S. (2002). Variance components models for gene–environment interaction in twin analysis. Twin Research, 5: 554571.Google Scholar
Purcell, S., Neale, B., Todd-Brown, K., Thomas, L., Ferreira, M. A., Bender, D., … & Sham, P. C. (2007). PLINK: a tool set for whole-genome association and population-based linkage analyses. American Journal of Human Genetics, 81: 559575.Google Scholar
Purcell, S. M., Wray, N. R., Stone, J. L., Visscher, P. M., O’Donovan, M. C., Sullivan, P. F., & Sklar, P. (2009). Common polygenic variation contributes to risk of schizophrenia and bipolar disorder. Nature, 460: 748752.Google Scholar
Scarr, S. & McCartney, K. (1983). How people make their own environments: a theory of genotype greater than environment effects. Child Development, 54: 424435.Google Scholar
Sham, P. C., Zhao, J. H., & Curtis, D. (1997). Optimal weighting scheme for affected sib-pair analysis of sibship data. Annals of Human Genetics, 61: 5967.Google Scholar
So, H. C., Li, M., & Sham, P. C. (2011). Uncovering the total heritability explained by all true susceptibility variants in a genome-wide association study. Genetic Epidemiology, 35: 447456.Google Scholar
Sullivan, P. F. (2010). The psychiatric GWAS consortium: big science comes to psychiatry. Neuron, 68: 182186.Google Scholar
Swagerman, S. C., Brouwer, R. M., de Geus, E. J. C., Hulshoff Pol, H. E., & Boomsma, D. I. (2014). Development and heritability of subcortical brain volumes at ages 9 and 12. Genes, Brain and Behavior, 13: 733742.Google Scholar
Towne, B., Almasy, L., Siervogel, R. M., & Blangero, J. (1999). Effects of genotype × sex interaction on linkage analysis of visual event-related evoked potentials. Genetic Epidemiology, 17: S355S360.Google Scholar
Van Beijsterveldt, C. E. M. & Boomsma, D. I. (1994). Genetics of the human electroencephalogram (EEG) and event-related brain potentials (ERPs): a review. Human Genetics, 94: 319330.Google Scholar
van der Sluis, S., Dolan, C. V., Neale, M. C., Boomsma, D. I., & Posthuma, D. (2006). Detecting genotype–environment interaction in monozygotic twin data: comparing the Jinks and Fulker test and a new test based on marginal maximum likelihood estimation. Twin Research and Human Genetics, 9: 377392.Google Scholar
Van Dongen, J., Jansen, R., Smit, D., Hottenga, J. J., Mbarek, H., Willemsen, G., … & de Geus, E. J. (2014). The contribution of the functional IL6R polymorphism rs2228145, eQTLs and other genome-wide SNPs to the heritability of plasma sIL-6R levels. Behavior Genetics, 44: 368382.Google Scholar
Van Dongen, J., Slagboom, P. E., Draisma, H. H., Martin, N. G., & Boomsma, D. I. (2012). The continuing value of twin studies in the omics era. Nature Reviews Genetics, 13: 640653.Google Scholar
Vink, J. M., Bartels, M., Van Beijsterveldt, T. C., Van Dongen, J., Van Beek, J. H., Distel, M. A., … & Boomsma, D. I. (2012). Sex differences in genetic architecture of complex phenotypes? PLoS One, 7: e47371.Google Scholar
Visscher, P. M., Brown, M. A., McCarthy, M. I., & Yang, J. (2012). Five years of GWAS discovery. American Journal of Human Genetics, 90: 724.Google Scholar
Visscher, P. M., Medland, S. E., Ferreira, M. A., Morley, K. I., Zhu, G., Cornes, B. K., … & Martin, N. G. (2006). Assumption-free estimation of heritability from genome-wide identity-by-descent sharing between full siblings. PLoS Genetics, 2: e41.Google Scholar
Visscher, P. M., Yang, J., & Goddard, M. E. (2010). A commentary on “common SNPs explain a large proportion of the heritability for human height” by Yang et al. (2010). Twin Research and Human Genetics, 13: 517524.Google Scholar
Wain, L. V., Verwoert, G. C., O’Reilly, P. F., Shi, G., Johnson, T., Johnson, A. D., … & vanDuijn, C. M. (2011). Genome-wide association study identifies six new loci influencing pulse pressure and mean arterial pressure. Nature Genetics, 43: 10051011.Google Scholar
Wang, B., Liao, C., Zhou, B., Cao, W., Lv, J., Yu, C., … & Li, L. (2015). Genetic contribution to the variance of blood pressure and heart rate: a systematic review and meta-regression of twin studies. Twin Research and Human Genetics, 18: 158170.Google Scholar
Willer, C. J., Li, Y., & Abecasis, G. R. (2010). METAL: fast and efficient meta-analysis of genomewide association scans. Bioinformatics, 26: 21902191.Google Scholar
Wray, N. R., Lee, S. H., Mehta, D., Vinkhuyzen, A. A., Dudbridge, F., & Middeldorp, C. M. (2014). Research review: polygenic methods and their application to psychiatric traits. Journal of Child Psychology and Psychiatry, 55: 10681087.Google Scholar
Wray, N. R., Middeldorp, C. M., Birley, A. J., Gordon, S. D., Sullivan, P. F., Visscher, P. M., … & Boomsma, D. I. (2008). Genome-wide linkage analysis of multiple measures of neuroticism of 2 large cohorts from Australia and the Netherlands. Archives of General Psychiatry, 65: 649658.Google Scholar
Wray, N. R., Yang, J., Hayes, B. J., Price, A. L., Goddard, M. E., & Visscher, P. M. (2013). Pitfalls of predicting complex traits from SNPs. Nature Reviews Genetics, 14: 507515.Google Scholar
Wu, T., Snieder, H., & de Geus, E. (2010). Genetic influences on cardiovascular stress reactivity. Neuroscience & Biobehavioral Reviews, 35: 5868.Google Scholar
Yang, J., Lee, S. H., Goddard, M. E., & Visscher, P. M. (2011). GCTA: a tool for genome-wide complex trait analysis. American Journal of Human Genetics, 88: 7682.Google Scholar
Yang, J., Zaitlen, N. A., Goddard, M. E., Visscher, P. M., & Price, A. L. (2014). Advantages and pitfalls in the application of mixed-model association methods. Nature Genetics, 46: 100106.Google Scholar
Zaitlen, N., Kraft, P., Patterson, N., Pasaniuc, B., Bhatia, G., Pollack, S., & Price, A. L. (2013). Using extended genealogy to estimate components of heritability for 23 quantitative and dichotomous traits. PLoS Genetics, 9: e1003520.Google Scholar
Zhu, Z., Bakshi, A., Vinkhuyzen, A. A., Hemani, G., Lee, S. H., Nolte, I. M., … & Yang, J. (2015). Dominance genetic variation contributes little to the missing heritability for human complex traits. American Journal of Human Genetics, 96: 377385.Google Scholar

References

Abbas, A. R., Baldwin, D., Ma, Y., Ouyang, W., Gurney, A., Martin, F., … & Clark, H. F. (2005). Immune response in silico (IRIS): immune-specific genes identified from a compendium of microarray expression data. Genes and Immunity, 6: 319331.Google Scholar
Alberts, B., Johnson, A., Lewis, J., Morgan, D., Raff, M., Roberts, K., & Walter, P. (2014). The Molecular Biology of the Cell. London: Garland.Google Scholar
Antoni, M. H., Lutgendorf, S. K., Blomberg, B., Stagl, J., Carver, C. S., Lechner, S., & Diaz, A. (2012). Transcriptional modulation of human leukocytes by cognitive-behavioral stress management in women undergoing treatment for breast cancer. Biological Psychiatry, 71: 366372.Google Scholar
Ashburner, M., Ball, C. A., Blake, J. A., Botstein, D., Butler, H., Cherry, J. M., … & Sherlock, G. (2000). Gene ontology: tool for the unification of biology. The Gene Ontology Consortium. Nature Genetics, 25: 2529.Google Scholar
Beissbarth, T. & Speed, T. P. (2004). GOstat: find statistically overrepresented gene ontologies within a group of genes. Bioinformatics, 20: 14641465.Google Scholar
Benjamini, Y. & Hochberg, Y. (1995). Controlling the false discovery rate: a practical and powerful approach to multiple testing. Journal of the Royal Statistical Society, Series B, 57: 289300.Google Scholar
Bergen, A. W., Mallick, A., Nishita, D., Wei, X., Michel, M., Wacholder, A., … & Andrews, J. A. (2012). Chronic psychosocial stressors and salivary biomarkers in emerging adults. Psychoneuroendocrinology, 37: 11581170.Google Scholar
Black, D. S., Cole, S. W., Irwin, M. R., Breen, E., St Cyr, N. M., Nazarian, N., … & Lavretsky, H. (2012). Yogic meditation reverses NF-kappaB and IRF-related transcriptome dynamics in leukocytes of family dementia caregivers in a randomized controlled trial. Psychoneuroendocrinology, 38: 348355.Google Scholar
Bolstad, B. M., Irizarry, R. A., Astrand, M., & Speed, T. P. (2003). A comparison of normalization methods for high density oligonucleotide array data based on variance and bias. Bioinformatics, 19: 185193.Google Scholar
Bower, J. E., Ganz, P. A., Irwin, M. R., Arevalo, J. M., & Cole, S. W. (2011). Fatigue and gene expression in human leukocytes: increased NF-kappaB and decreased glucocorticoid signaling in breast cancer survivors with persistent fatigue. Brain, Behavior, and Immunity, 25: 147150.Google Scholar
Bower, J. E., Greendale, G., Crosswell, A. D., Garet, D., Sternlieb, B., Ganz, P. A., … & Cole, S. W. (2014). Yoga reduces inflammatory signaling in fatigued breast cancer survivors: a randomized controlled trial. Psychoneuroendocrinology, 43: 2029.Google Scholar
Cao, J. & Zhang, S. (2014). Multiple comparison procedures. Journal of the American Medical Association, 312: 543544.Google Scholar
Capitanio, J. P. & Cole, S. W. (2015). Social instability and immunity in rhesus monkeys: the role of the sympathetic nervous system. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 370. doi: 10.1098/rstb.2014.0104Google Scholar
Carey, M. & Smale, S. T. (2001). Transcriptional Regulation in Eukaryotes: Concepts, Strategies, and Techniques. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press.Google Scholar
Chikina, M., Zaslavsky, E., & Sealfon, S. C. (2015). CellCODE: a robust latent variable approach to differential expression analysis for heterogeneous cell populations. Bioinformatics, 31: 15841591.Google Scholar
Clark, D. P. & Russell, L. D. (2005). Molecular Biology: Made Simple and Fun. St. Louis, MO: Cache River Press.Google Scholar
Cole, S. W. (2010). Elevating the perspective on human stress genomics. Psychoneuroendocrinology, 35: 955962.Google Scholar
Cole, S. W. (2013). Social regulation of human gene expression: mechanisms and implications for public health. American Journal of Public Health, 103: S84S92.Google Scholar
Cole, S. W. (2014). Human social genomics. PLoS Genetics, 10: e1004601.Google Scholar
Cole, S. W., Arevalo, J. M., Ruggerio, A. M., Heckman, J. J., & Suomi, S. (2012). Transcriptional modulation of the developing immune system by early life social adversity. Proceedings of the National Academy of Sciences of the USA, 109: 2057820583.Google Scholar
Cole, S. W., Arevalo, J., Takahashi, R., Sloan, E. K., Lutgendorf, S., Sood, A. K., … & Seeman, T. E. (2010). Computational identification of gene–social environment interaction at the human IL6 locus. Proceedings of the National Academy of Sciences of the USA, 107: 56815686.Google Scholar
Cole, S. W., Capitanio, J. P., Chun, K., Arevalo, J. M. G., Ma, J., & Cacioppo, J. T. (2015a). Myeloid differentiation architecture of leukocyte transcriptome dynamics in perceived social isolation. Proceedings of the National Academy of Sciences of the USA, 112: 1514215147.Google Scholar
Cole, S. W., Galic, Z., & Zack, J. A. (2003). Controlling false-negative errors in microarray differential expression analysis: a PRIM approach. Bioinformatics, 19: 18081816.Google Scholar
Cole, S. W., Hawkley, L. C., Arevalo, J. M., & Cacioppo, J. T. (2011). Transcript origin analysis identifies antigen-presenting cells as primary targets of socially regulated gene expression in leukocytes. Proceedings of the National Academy of Sciences of the USA, 108: 30803085.Google Scholar
Cole, S. W., Hawkley, L. C., Arevalo, J. M., Sung, C. Y., Rose, R. M., & Cacioppo, J. T. (2007). Social regulation of gene expression in human leukocytes. Genome Biology, 8: 113.Google Scholar
Cole, S. W., Levine, M. E., Arevalo, J. M., Ma, J., Weir, D. R., & Crimmins, E. M. (2015b). Loneliness, eudaimonia, and the human conserved transcriptional response to adversity. Psychoneuroendocrinology, 62: 1117.Google Scholar
Cole, S. W., Mendoza, S. P., & Capitanio, J. P. (2009). Social stress desensitizes lymphocytes to regulation by endogenous glucocorticoids: insights from in vivo cell trafficking dynamics in rhesus macaques. Psychosomatic Medicine, 71: 591597.Google Scholar
Cole, S. W., Yan, W., Galic, Z., Arevalo, J., & Zack, J. A. (2005). Expression-based monitoring of transcription factor activity: the TELiS database. Bioinformatics, 21: 803810.Google Scholar
Collado-Hidalgo, A., Sung, C., & Cole, S. (2006). Adrenergic inhibition of innate anti-viral response: PKA blockade of Type I interferon gene transcription mediates catecholamine support for HIV-1 replication. Brain, Behavior, and Immunity, 20: 552563.Google Scholar
Creswell, J. D., Irwin, M. R., Burklund, L. J., Lieberman, M. D., Arevalo, J. M., Ma, J., … & Cole, S. W.(2012). Mindfulness-based stress reduction training reduces loneliness and pro-inflammatory gene expression in older adults: a small randomized controlled trial. Brain, Behavior, and Immunity, 26: 10951101.Google Scholar
Dale, D. C., Fauci, A. S., Guerry, D. I., & Wolff, S. M. (1975). Comparison of agents producing a neutrophilic leukocytosis in man: hydrocortisone, prednisone, endotoxin, and etiocholanolone. Journal of Clinical Investigation, 56: 808813.Google Scholar
Editors (2014). Honing our reading skills. Nature Biotechnology, 32: 845.Google Scholar
Efron, B. & Tibshirani, R. J. (1993). An Introduction to the Bootstrap. New York: Chapman & Hall.Google Scholar
Eisen, M. B., Spellman, P. T., Brown, P. O., & Botstein, D. (1998). Cluster analysis and display of genome-wide expression patterns. Proceedings of the National Academy of Sciences of the USA, 95: 1486314868.Google Scholar
Eisenberger, N. I., Inagaki, T. K., Mashal, N. M., & Irwin, M. R. (2010). Inflammation and social experience: an inflammatory challenge induces feelings of social disconnection in addition to depressed mood. Brain, Behavior, and Immunity, 24: 558563.Google Scholar
Ewens, W. J. & Grant, G. R. (2005). Statistical Methods in Bioinformatics. New York: Springer.Google Scholar
Fauci, A. S. & Dale, D. C. (1975). The effect of hydrocortisone on the kinetics of normal human lymphocytes. Blood, 46: 235243.Google Scholar
Fauci, A. S., Dale, D. C., & Balow, J. E. (1976). Glucocorticosteroid therapy: mechanisms of action and clinical considerations. Annals of Internal Medicine, 84: 304315.Google Scholar
Felger, J. C., Cole, S. W., Pace, T. W., Hu, F., Woolwine, B. J., Doho, G. H., … & Miller, A. H. (2012). Molecular signatures of peripheral blood mononuclear cells during chronic interferon-alpha treatment: relationship with depression and fatigue. Psychological Medicine, 42: 15911603.Google Scholar
Fredrickson, B. L., Grewen, K. M., Algoe, S. B., Firestine, A. M., Arevalo, J. M. G., Ma, J., & Cole, S. W. (2015). Psychological well-being and the human conserved transcriptional response to adversity. PLoS One, 10: e0121839.Google Scholar
Fredrickson, B. L., Grewen, K. M., Coffey, K. A., Algoe, S. B., Firestine, A. M., Arevalo, J. M., … & Cole, S. W. (2013). A functional genomic perspective on human well-being. Proceedings of the National Academy of Sciences of the USA, 110: 1368413689.Google Scholar
Gaujoux, R. & Seoighe, C. (2013). CellMix: a comprehensive toolbox for gene expression deconvolution. Bioinformatics, 29: 22112212.Google Scholar
Gibson, G. (2008). The environmental contribution to gene expression profiles. Nature Reviews Genetics, 9: 575581.Google Scholar
Goldinger, A., Henders, A. K., McRae, A. F., Martin, N. G., Gibson, G., Montgomery, G. W., … & Powell, J. E. (2013). Genetic and nongenetic variation revealed for the principal components of human gene expression. Genetics, 195: 11171128.Google Scholar
Gray, J. D., Rubin, T. G., Hunter, R. G., & McEwen, B. S. (2014). Hippocampal gene expression changes underlying stress sensitization and recovery. Molecular Psychiatry, 19: 11711178.Google Scholar
Hastie, T., Tibshirani, R., & Friedman, J. (2001). The Elements of Statistical Learning. New York: Springer.Google Scholar
Heidt, T., Sager, H. B., Courties, G., Dutta, P., Iwamoto, Y., Zaltsman, A., … & Nahrendorf, M. (2014). Chronic variable stress activates hematopoietic stem cells. Nature Medicine, 20: 754758.Google Scholar
Huang da, W., Sherman, B. T., & Lempicki, R. A. (2009). Bioinformatics enrichment tools: paths toward the comprehensive functional analysis of large gene lists. Nucleic Acids Research, 37: 113.Google Scholar
Irwin, M., Olmstead, R., Breen, E., Witarama, T., Carrillo, C., Sadeghi, N., … & Cole, S. (2014). Tai Chi, cellular inflammation, and transcriptome dynamics in breast cancer survivors with insomnia: a randomized controlled trial. Journal of the National Cancer Institute, 50: 295301.Google Scholar
Karssen, A. M., Her, S., Li, J. Z., Patel, P. D., Meng, F., Bunney, W. E. Jr., … & Lyons, D. M. (2007). Stress-induced changes in primate prefrontal profiles of gene expression. Molecular Psychiatry, 12: 10891102.Google Scholar
Kramer, A., Green, J., Pollard, J. Jr., & Tugendreich, S. (2014). Causal analysis approaches in Ingenuity Pathway Analysis. Bioinformatics, 30: 523530.Google Scholar
Kutner, M., Nachtsheim, C., Neter, J., & Li, W. (2004). Applied Linear Statistical Models. New York: McGraw-Hill/Irwin.Google Scholar
Lahdesmaki, H., Shmulevich, L., Dunmire, V., Yli-Harja, O., & Zhang, W. (2005). In silico microdissection of microarray data from heterogeneous cell populations. BMC Bioinformatics, 6: 54.Google Scholar
Landmark-Hoyvik, H., Reinertsen, K. V., Loge, J. H., Fossa, S. D., Borresen-Dale, A. L., & Dumeaux, V. (2009). Alterations of gene expression in blood cells associated with chronic fatigue in breast cancer survivors. Pharmacogenomics Journal, 9: 333340.Google Scholar
Langfelder, P. & Horvath, S. (2008). WGCNA: an R package for weighted correlation network analysis. BMC Bioinformatics, 9: 559.Google Scholar
Li, S., Labaj, P. P., Zumbo, P., Sykacek, P., Shi, W., Shi, L., … & Mason, C. E. (2014a). Detecting and correcting systematic variation in large-scale RNA sequencing data. Nature Biotechnology, 32: 888895.Google Scholar
Li, S., Tighe, S. W., Nicolet, C. M., Grove, D., Levy, S., Farmerie, W., … & Mason, C. E. (2014b). Multi-platform assessment of transcriptome profiling using RNA-seq in the ABRF next-generation sequencing study. Nature Biotechnology, 32: 915925.Google Scholar
Livak, K. J. & Schmittgen, T. D. (2001). Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) method. Methods, 25: 402408.Google Scholar
Lutgendorf, S. K., Degeest, K., Sung, C. Y., Arevalo, J. M., Penedo, F., Lucci, J. 3rd, … & Cole, S. W. (2009). Depression, social support, and beta-adrenergic transcription control in human ovarian cancer. Brain, Behavior, and Immunity, 23: 176183.Google Scholar
Malone, J. H. & Oliver, B. (2011). Microarrays, deep sequencing and the true measure of the transcriptome. BMC Biology, 9: 34.Google Scholar
McCulloch, C. E., Searle, S. R., & Neuhaus, J. M. (2008). Generalized, Linear, and Mixed Models. Hoboken, NJ: John Wiley.Google Scholar
Miller, G., Chen, E., & Cole, S. W. (2009). Health psychology: developing biologically plausible models linking the social world and physical health. Annual Review of Psychology, 60: 501524.Google Scholar
Miller, G. E., Chen, E., Sze, J., Marin, T., Arevalo, J. M., Doll, R., … & Cole, S. W. (2008). A functional genomic fingerprint of chronic stress in humans: blunted glucocorticoid and increased NF-kappaB signaling. Biological Psychiatry, 64, 266272.Google Scholar
Miller, G. E. & Cole, S. W. (2010). Functional genomic approaches in behavioral medicine research. In Steptoe, A. (ed.), Handbook of Behavioral Medicine: Methods and Applications (pp. 443454). New York: Springer.Google Scholar
Miller, G. E., Murphy, M. L. M., Cashman, R., Ma, R., Ma, J., Arevalo, J. M. G., … & Cole, S. W. (2014). Greater inflammatory activity and blunted glucocorticoid signaling in monocytes of chronically stressed caregivers. Brain, Behavior and Immunity, 41: 191199.Google Scholar
Miller, R. G. (1986). Beyond ANOVA: Basics of Applied Statistics. New York: John Wiley.Google Scholar
Munro, S. A., Lund, S. P., Pine, P. S., Binder, H., Clevert, D. A., Conesa, A., … & Salit, M. (2014). Assessing technical performance in differential gene expression experiments with external spike-in RNA control ratio mixtures. Nature Communications, 5: 5125.Google Scholar
Nath, A. P., Arafat, D., & Gibson, G. (2012). Using blood informative transcripts in geographical genomics: impact of lifestyle on gene expression in Fijians. Frontiers in Genetics, 3: 243.Google Scholar
Norris, A. W. & Kahn, C. R. (2006). Analysis of gene expression in pathophysiological states: balancing false discovery and false negative rates. Proceedings of the National Academy of Sciences of the USA., 103: 649653.Google Scholar
O’Donovan, A., Sun, B., Cole, S., Rempel, H., Lenoci, M., Pulliam, L., & Neylan, T. (2011). Transcriptional control of monocyte gene expression in post-traumatic stress disorder. Disease Markers, 30: 123132.Google Scholar
Powell, N. D., Sloan, E. K., Bailey, M. T., Arevalo, J. M., Miller, G. E., Chen, E., … & Cole, S. W. (2013). Social stress up-regulates inflammatory gene expression in the leukocyte transcriptome via beta-adrenergic induction of myelopoiesis. Proceedings of the National Academy of Sciences of the USA, 110: 1657416579.Google Scholar
Preininger, M., Arafat, D., Kim, J., Nath, A. P., Idaghdour, Y., Brigham, K. L., & Gibson, G. (2013). Blood-informative transcripts define nine common axes of peripheral blood gene expression. PLoS Genetics, 9: e1003362.Google Scholar
Proud, D., Turner, R. B., Winther, B., Wiehler, S., Tiesman, J. P., Reichling, T. D., … & Clymer, J. W. (2008). Gene expression profiles during in vivo human rhinovirus infection: insights into the host response. American Journal of Respiratory and Critical Care Medicine, 178: 962968.Google Scholar
QIAGEN (2015). Ingenuity Pathway Analysis.Google Scholar
Ramilo, O., Allman, W., Chung, W., Mejias, A., Ardura, M., Glaser, C., … & Chaussabel, D. (2007). Gene expression patterns in blood leukocytes discriminate patients with acute infections. Blood, 109: 20662077.Google Scholar
Richlin, V. A., Arevalo, J. M., Zack, J. A., & Cole, S. W. (2004). Stress-induced enhancement of NF-kappaB DNA-binding in the peripheral blood leukocyte pool: effects of lymphocyte redistribution. Brain, Behavior, and Immunity, 18: 231237.Google Scholar
Risso, D., Ngai, J., Speed, T. P., & Dudoit, S. (2014). Normalization of RNA-seq data using factor analysis of control genes or samples. Nature Biotechnology, 32: 896902.Google Scholar
SEQC/MAQC-III Consortium (2014). A comprehensive assessment of RNA-seq accuracy, reproducibility and information content by the Sequencing Quality Control Consortium. Nature Biotechnology, 32: 903914.Google Scholar
Shen-Orr, S. S. & Gaujoux, R. (2013). Computational deconvolution: extracting cell type-specific information from heterogeneous samples. Current Opinion in Immunology, 25: 571578.Google Scholar
Shen-Orr, S. S., Tibshirani, R., Khatri, P., Bodian, D. L., Staedtler, F., Perry, N. M., … & Butte, A. J. (2010). Cell type-specific gene expression differences in complex tissues. Nature Methods, 7: 287289.Google Scholar
Shi, L., Jones, W. D., Jensen, R. V., Harris, S. C., Perkins, R. G., Goodsaid, F. M., … & Tong, W. (2008). The balance of reproducibility, sensitivity, and specificity of lists of differentially expressed genes in microarray studies. BMC Bioinformatics, 9: S10.Google Scholar
Slavich, G. M. & Cole, S. W. (2013). The emerging field of human social genomics. Clinical Psychological Science, 1: 331348.Google Scholar
Sloan, E. K., Capitanio, J. P., & Cole, S. W. (2008). Stress-induced remodeling of lymphoid innervation. Brain, Behavior, and Immunity, 22: 1521.Google Scholar
Sloan, E. K., Capitanio, J. P., Tarara, R. P., Mendoza, S. P., Mason, W. A., & Cole, S. W. (2007). Social stress enhances sympathetic innervation of primate lymph nodes: mechanisms and implications for viral pathogenesis. Journal of Neuroscience, 27: 88578865.Google Scholar
Stone, E. A. & Ayroles, J. F. (2009). Modulated modularity clustering as an exploratory tool for functional genomic inference. PLoS Genetics, 5: e1000479.Google Scholar
Strachan, T. & Read, A. P. (2004). Human Molecular Genetics 3. London: Garland.Google Scholar
Su, A. I., Wiltshire, T., Batalov, S., Lapp, H., Ching, K. A., Block, D., … & Hogenesch, J. B. (2004). A gene atlas of the mouse and human protein-encoding transcriptomes. Proceedings of the National Academy of Sciences of the USA, 101: 60626067.Google Scholar
Su, Z., Fang, H., Hong, H., Shi, L., Zhang, W., Zhang, Y., … & Tong, W. (2014). An investigation of biomarkers derived from legacy microarray data for their utility in the RNA-seq era. Genome Biology, 15: 523.Google Scholar
Subramanian, A., Tamayo, P., Mootha, V. K., Mukherjee, S., Ebert, B. L., Gillette, M. A., … & Mesirov, P. (2005). Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proceedings of the National Academy of Sciences of the USA, 102: 1554515550.Google Scholar
Tabassum, R., Nath, A., Preininger, M., & Gibson, G. (2013). Geographical, environmental and pathophysiological influences on the human blood transcriptome. Current Genetic Medicine Reports, 1: 203211.Google Scholar
Tung, J., Barreiro, L. B., Johnson, Z. P., Hansen, K. D., Michopoulos, V., Toufexis, D., … & Gilad, Y. (2012). Social environment is associated with gene regulatory variation in the rhesus macaque immune system. Proceedings of the National Academy of Sciences of the USA, 109: 64906495.Google Scholar
Tusher, V. G., Tibshirani, R., & Chu, G. (2001). Significance analysis of microarrays applied to the ionizing radiation response. Procedings of the National Academy of Sciences of the USA, 98: 51165121.Google Scholar
Wang, C., Gong, B., Bushel, P. R., Thierry-Mieg, J., Thierry-Mieg, D., Xu, J., … & Tong, W. (2014). The concordance between RNA-seq and microarray data depends on chemical treatment and transcript abundance. Nature Biotechnology, 32: 926932.Google Scholar
Wingo, A. P. & Gibson, G. (2015). Blood gene expression profiles suggest altered immune function associated with symptoms of generalized anxiety disorder. Brain, Behavior, and Immunity, 43: 184191.Google Scholar
Witten, D. M. & Tibshirani, R. (2007). A comparison of fold-change and the T-statistic for microarray data analysis. Stanford University.Google Scholar
Yu, J., Cliften, P. F., Juehne, T. I., Sinnwell, T. M., Sawyer, C. S., Sharma, M., … & Head, R. D. (2015). Multi-platform assessment of transcriptional profiling technologies utilizing a precise probe mapping methodology. BMC Genomics, 16: 710.Google Scholar
Zhang, W., Yu, Y., Hertwig, F., Thierry-Mieg, J., Thierry-Mieg, D., Wang, J., … & Fischer, M. (2015). Comparison of RNA-seq and microarray-based models for clinical endpoint prediction. Genome Biology, 16: 133.Google Scholar

References

Amit, I., Garber, M., Chevrier, N., Leite, A. P., Donner, Y., Eisenhaure, T., … & Regev, A. (2009). Unbiased reconstruction of a mammalian transcriptional network mediating pathogen responses. Science, 326: 257263.Google Scholar
Antoni, M. H. (2013). Psychosocial intervention effects on adaptation, disease course and biobehavioral processes in cancer. Brain, Behavior, and Immunity, 30: S88S98.Google Scholar
Aschbacher, K., Epel, E., Wolkowitz, O. M., Prather, A. A., Puterman, E., & Dhabhar, F. S. (2012). Maintenance of a positive outlook during acute stress protects against pro-inflammatory reactivity and future depressive symptoms. Brain, Behavior, and Immunity, 26: 346352.Google Scholar
Avitsur, R., Kinsey, S. G., Bidor, K., Bailey, M. T., Padgett, D. A., & Sheridan, J. F. (2007). Subordinate social status modulates the vulnerability to the immunological effects of social stress. Psychoneuroendocrinology, 32: 10971105.Google Scholar
Barton, G. M. (2008). A calculated response: control of inflammation by the innate immune system. Journal of Clinical Investigation, 118: 413420.Google Scholar
Benedetti, F., Lucca, A., Brambilla, F., Colombo, C., & Smeraldi, E. (2002). Interleukine-6 serum levels correlate with response to antidepressant sleep deprivation and sleep phase advance. Progress in Neuro-Psychopharmacology and Biological Psychiatry, 26: 11671170.Google Scholar
Berkenbosch, F., van Oers, J., del Rey, A., Tilders, F., & Besedovsky, H. (1987). Corticotropin-releasing factor-producing neurons in the rat activated by interleukin-1. Science, 238: 524526.Google Scholar
Besedovsky, H., del Rey, A., Sorkin, E., & Dinarello, C. A. (1986). Immunoregulatory feedback between interleukin-1 and glucocorticoid hormones. Science, 233: 5254.Google Scholar
Bierhaus, A., Wolf, J., Andrassy, M., Rohleder, N., Humpert, P. M., Petrov, D., … & Nawroth, P. P. (2003). A mechanism converting psychosocial stress into mononuclear cell activation. Proceedings of the National Academy of Sciences of the USA, 100: 19201925.Google Scholar
Bower, J. E., Ganz, P. A., Aziz, N., Olmstead, R., Irwin, M. R., & Cole, S. W. (2007). Inflammatory responses to psychological stress in fatigued breast cancer survivors: relationship to glucocorticoids. Brain, Behavior and Immunity, 21: 251258.Google Scholar
Bower, J. E. & Irwin, M. R. (2016). Mind–body therapies and control of inflammatory biology: a descriptive review. Brain, Behavior, and Immunity, 51: 111.Google Scholar
Breen, E. C., Rezai, A. I., Nakajima, K., Beall, G. N., Mitsuyasu, R. T., Hirano, T., … & Martinez-Maza, O. (1990). Infection with HIV is associated with elevated IL-6 levels and production. Journal of Immunology, 144: 480484.Google Scholar
Brown, R., Pang, G., Husband, A. J., & King, M. G. (1989). Suppression of immunity to influenza virus infection in the respiratory tract following sleep disturbance. Regional Immunology, 2: 321325.Google Scholar
Bull, S. J., Huezo-Diaz, P., Binder, E. B., Cubells, J. F., Ranjith, G., Maddock, G., … & Pariante, C. M. (2009). Functional polymorphisms in the interleukin-6 and serotonin transporter genes, and depression and fatigue induced by interferon-alpha and ribavirin treatment. Molecular Psychiatry, 14: 10951104.Google Scholar
Cappuccio, F. P., D’Elia, L., Strazzullo, P., & Miller, M. A. (2010). Sleep duration and all-cause mortality: a systematic review and meta-analysis of prospective studies. Sleep, 33: 585592.Google Scholar
Carroll, J. E., Low, C. A., Prather, A. A., Cohen, S., Fury, J. M., Ross, D. C., & Marsland, A. L. (2011). Negative affective responses to a speech task predict changes in interleukin (IL)-6. Brain, Behavior, and Immunity, 25: 232238.Google Scholar
Chen, E., Miller, G. E., Walker, H. A., Arevalo, J. M., Sung, C. Y., & Cole, S. W. (2009). Genome-wide transcriptional profiling linked to social class in asthma. Thorax, 64: 3843.Google Scholar
Cohen, S., Doyle, W. J., Alper, C. M., Janicki-Deverts, D., & Turner, R. B. (2009). Sleep habits and susceptibility to the common cold. Archives of Internal Medicine, 169: 6267.Google Scholar
Cohen, S., Janicki-Deverts, D., & Miller, G. E. (2007). Psychological stress and disease. Journal of the American Medical Association, 298: 16851687.Google Scholar
Cohen, S., Tyrrell, D. A. J., & Smith, A. P. (1991). Psychological stress and susceptibility to the common cold. New England Journal of Medicine, 325: 606612.Google Scholar
Cole, S. W. (2009). Chronic inflammation and breast cancer recurrence. Journal of Clinical Oncology, 27: 34183419.Google Scholar
Cole, S. W., Arevalo, J. M., Manu, K., Telzer, E. H., Kiang, L., Bower, J. E., … & Fuligni, A. J. (2011a). Antagonistic pleiotropy at the human IL6 promoter confers genetic resilience to the pro-inflammatory effects of adverse social conditions in adolescence. Developmental Psychology, 47: 11731180.Google Scholar
Cole, S. W., Arevalo, J. M., Takahashi, R., Sloan, E. K., Lutgendorf, S. K., Sood, A. K., … & Seeman, T. E. (2010). Computational identification of gene–social environment interaction at the human IL6 locus. Proceedings of the National Academy of Sciences of the USA, 107: 56815686.Google Scholar
Cole, S. W., Hawkley, L. C., Arevalo, J. M., & Cacioppo, J. T. (2011b). Transcript origin analysis identifies antigen-presenting cells as primary targets of socially regulated gene expression in leukocytes. Proceedings of the National Academy of Sciences of the USA, 108: 30803085.Google Scholar
Cole, S. W., Hawkley, L. C., Arevalo, J. M., Sung, C. Y., Rose, R. M., & Cacioppo, J. T. (2007). Social regulation of gene expression in human leukocytes. Genome Biology, 8: 113.Google Scholar
Cole, S. W. & Kemeny, M. E. (1997). Psychobiology of HIV infection. Critical Reviews in Neurobiology, 11: 289321.Google Scholar
Cole, S. W., Kemeny, M. E., & Taylor, S. E. (1997). Social identity and physical health: accelerated HIV progression in rejection-sensitive gay men. Journal of Personality and Social Psychology, 72: 320335.Google Scholar
Cole, S. W., Korin, Y. D., Fahey, J. L., & Zack, J. A. (1998). Norepinephrine accelerates HIV replication via protein kinase A-dependent effects on cytokine production. Journal of Immunology, 161: 610616.Google Scholar
Collado-Hidalgo, A., Bower, J. E., Ganz, P. A., Irwin, M. R., & Cole, S. W. (2008). Cytokine gene polymorphisms and fatigue in breast cancer survivors: early findings. Brain, Behavior and Immunology, 22: 11971200.Google Scholar
Collado-Hidalgo, A., Sung, C., & Cole, S. (2006). Adrenergic inhibition of innate anti-viral response: PKA blockade of Type I interferon gene transcription mediates catecholamine support for HIV-1 replication. Brain, Behavior and Immunology, 20: 552563.Google Scholar
Creswell, J. D., Irwin, M. R., Burklund, L. J., Lieberman, M. D., Arevalo, J. M., Ma, J., … & Cole, S. W. (2012). Mindfulness-based stress reduction training reduces loneliness and pro-inflammatory gene expression in older adults: a small randomized controlled trial. Brain, Behavior and Immunology, 26: 10951101.Google Scholar
Curfs, J. H., Meis, J. F., & Hoogkamp-Korstanje, J. A. (1997). A primer on cytokines: sources, receptors, effects, and inducers. Clinical Microbiology Reviews, 10: 742780.Google Scholar
Danese, A., Moffitt, T. E., Pariante, C. M., Ambler, A., Poulton, R., & Caspi, A. (2008). Elevated inflammation levels in depressed adults with a history of childhood maltreatment. Archives of Generaal Psychiatry, 65: 409415.Google Scholar
Dantzer, R., O’Connor, J. C., Freund, G. G., Johnson, R. W., & Kelley, K. W. (2008). From inflammation to sickness and depression: when the immune system subjugates the brain. Nature Reviews Neuroscience, 9: 4656.Google Scholar
Davis, M. C., Zautra, A. J., Younger, J., Motivala, S. J., Attrep, J., & Irwin, M. R. (2008). Chronic stress and regulation of cellular markers of inflammation in rheumatoid arthritis. Brain, Behavior, and Immunity, 22: 2432.Google Scholar
DellaGioia, N. & Hannestad, J. (2010). A critical review of human endotoxin administration as an experimental paradigm of depression. Neuroscience & Biobehavioral Reviews, 34: 130143.Google Scholar
Dew, M. A., Hoch, C. C., Buysse, D. J., Monk, T. H., Begley, A. E., Houck, P. R., … & Reynolds, C. F. III (2003). Healthy older adults’ sleep predicts all-cause mortality at 4 to 19 years of follow-up. Psychosomatic Medicine, 65: 6373.Google Scholar
Dhabhar, F. S., Malarkey, W. B., Neri, E., & McEwen, B. S. (2012). Stress-induced redistribution of immune cells–from barracks to boulevards to battlefields: a tale of three hormones – Curt Richter Award winner. Psychoneuroendocrinology, 37: 13451368.Google Scholar
Dickerson, S. S., Gable, S. L., Irwin, M. R., Aziz, N., & Kemeny, M. E. (2009). Social-evaluative threat and proinflammatory cytokine regulation: an experimental laboratory investigation. Psychological Science, 20: 12371244.Google Scholar
Dickerson, S. S. & Kemeny, M. E. (2004). Acute stressors and cortisol responses: a theoretical integration and synthesis of laboratory research. Psychological Bulletin, 130: 355391.Google Scholar
Dickerson, S. S., Kemeny, M. E., Aziz, N., Kim, K. H., & Fahey, J. L. (2004). Immunological effects of induced shame and guilt. Psychosomtic Medicine, 66: 124131.Google Scholar
Edwards, K. M., Burns, V. E., Reynolds, T., Carroll, D., Drayson, M., & Ring, C. (2006). Acute stress exposure prior to influenza vaccination enhances antibody response in women. Brain, Behavior, and Immunology, 20: 159168.Google Scholar
Eisenberger, N. I., Berkman, E. T., Inagaki, T. K., Rameson, L. T., Mashal, N. M., & Irwin, M. R. (2010a). Inflammation-induced anhedonia: endotoxin reduces ventral striatum responses to reward. Biological Psychiatry, 68: 748754.Google Scholar
Eisenberger, N. I., Inagaki, T. K., Mashal, N. M., & Irwin, M. R. (2010b). Inflammation and social experience: an inflammatory challenge induces feelings of social disconnection in addition to depressed mood. Brain, Behavior, and Immunity, 24: 558563.Google Scholar
Eisenberger, N. I., Inagaki, T. K., Rameson, L. T., Mashal, N. M., & Irwin, M. R. (2009). An fMRI study of cytokine-induced depressed mood and social pain: the role of sex differences. NeuroImage, 47: 881890.Google Scholar
Engler, H., Bailey, M. T., Engler, A., & Sheridan, J. F. (2004). Effects of repeated social stress on leukocyte distribution in bone marrow, peripheral blood and spleen. Journal of Neuroimmunology, 148: 106115.Google Scholar
Frank, M. G., Hendricks, S. E., Johnson, D., & Burke, W. J. (1999). Antidepressants augment natural killer cell activity: in vivo and in vitro. Neuropsychobiology, 39: 1824.Google Scholar
Friedman, E. M. & Irwin, M. (1997). Modulation of immune cell function by the autonomic nervous system. Pharmacology Therapy, 74: 2738.Google Scholar
Fuligni, A. J., Telzer, E. H., Bower, J., Cole, S. W., Kiang, L., & Irwin, M. R. (2009a). A preliminary study of daily interpersonal stress and C-reactive protein levels among adolescents from Latin American and European backgrounds. Psychosomatic Medicine, 71 (3): 329–33.Google Scholar
Fuligni, A. J., Telzer, E. H., Bower, J., Irwin, M. R., Kiang, L., & Cole, S. W. (2009b). Daily family assistance and inflammation among adolescents from Latin American and European backgrounds. Brain, Behavior, and Immunity, 23: 803809.Google Scholar
Gimeno, D., Kivimäki, M., Brunner, E. J., Elovainio, M., De Vogli, R., Steptoe, A., … & Ferrie, J. E. (2009). Associations of C-reactive protein and interleukin-6 with cognitive symptoms of depression: 12-year follow-up of the Whitehall II study. Psychological Medicine, 39: 413423.Google Scholar
Glaser, R. & Kiecolt-Glaser, J. K. (2005). Stress-induced immune dysfunction: implications for health. Nature Reviews Immunology, 5: 243251.Google Scholar
Goebel, M. U., Mills, P. J., Irwin, M. R., & Ziegler, M. G. (2000). Interleukin-6 and tumor necrosis factor-alpha production after acute psychological stress, exercise, and infused isoproterenol: differential effects and pathways. Psychosomatic Medicine, 62: 591598.Google Scholar
Grebe, K. M., Takeda, K., Hickman, H. D., Bailey, A. L., Embry, A. C., Bennink, J. R., & Yewdell, J. W. (2009). Cutting edge: sympathetic nervous system increases proinflammatory cytokines and exacerbates influenza A virus pathogenesis. Journal of Immunology, 184: 540544.Google Scholar
Gruys, E., Toussaint, M. J. M., Niewold, T. A., & Koopmans, S. J. (2005). Acute phase reaction and acute phase proteins. Journal of Zhejiang University Science B, 6: 10451056.Google Scholar
Haack, M., Sanchez, E., & Mullington, J. M. (2007). Elevated inflammatory markers in response to prolonged sleep restriction are associated with increased pain experience in healthy volunteers. Sleep, 30: 11451152.Google Scholar
Harrison, N. A., Brydon, L., Walker, C., Gray, M. A., Steptoe, A., & Critchley, H. D. (2009). Inflammation causes mood changes through alterations in subgenual cingulate activity and mesolimbic connectivity. Biological Psychiatry, 66: 407414.Google Scholar
Hart, B. L. (1988). Biological basis of the behavior of sick animals. Neuroscience & Biobehavioral Reviews, 12: 123137.Google Scholar
Howren, M. B., Lamkin, D. M., & Suls, J. (2009). Associations of depression with C-reactive protein, IL-1, and IL-6: a meta-analysis. Psychosomatic Medicine, 71: 171186.Google Scholar
Hubbard, A. K. & Rothlein, R. (2000). Intercellular adhesion molecule-1 (ICAM-1) expression and cell signaling cascades. Free Radical Biology & Medicine, 28: 13791386.Google Scholar
Imeri, L. & Opp, M. R. (2009). How (and why) the immune system makes us sleep. Nature Reviews Neuroscience, 10: 199210.Google Scholar
Irwin, M. R. (2015). Why sleep is important for health: a psychoneuroimmunology perspective. Annual Review of Psychology, 66: 2.12.30.Google Scholar
Irwin, M. R., Carrillo, C., & Olmstead, R. (2010). Sleep loss activates cellular markers of inflammation: sex differences. Brain, Behavior, and Immunity, 24: 5457.Google Scholar
Irwin, M. R. & Cole, S. W. (2011). Reciprocal regulation of the neural and innate immune systems. Nature Reviews Immunology, 11: 625632.Google Scholar
Irwin, M. R., Costlow, C., Williams, H., Artin, K. H., Chan, C. Y., Stinson, D. L., … & Oxman, M. N. (1998). Cellular immunity to varicella-zoster virus in patients with major depression. Journal of Infectious Diseases, 178: S104S108.Google Scholar
Irwin, M. R., Lacher, U., & Caldwell, C. (1992). Depression and reduced natural killer cytotoxicity: a longitudinal study of depressed patients and control subjects. Psychological Medicine, 22: 10451050.Google Scholar
Irwin, M. R., Levin, M. J., Carrillo, C., Olmstead, R., Lucko, A., Lang, N., … & Oxman, M. N. (2011). Major depressive disorder and immunity to varicella-zoster virus in the elderly. Brain, Behavior, and Immunity, 25: 759766.Google Scholar
Irwin, M. R., Levin, M. J., Laudenslager, M. L., Olmstead, R., Lucko, A., Lang, N., … & Oxman, M. N. (2013). Varicella zoster virus-specific immune responses to a herpes zoster vaccine in elderly recipients with major depression and the impact of antidepressant medications. Clinical Infectious Diseases, 56: 10851093.Google Scholar
Irwin, M. R., McClintick, J., Costlow, C., Fortner, M., White, J., & Gillin, J. C. (1996). Partial night sleep deprivation reduces natural killer and cellular immune responses in humans. FASEB Journal, 10: 643653.Google Scholar
Irwin, M. R. & Miller, A. H. (2007). Depressive disorders and immunity: 20 years of progress and discovery. Brain, Behavior, and Immunity, 21: 374383.Google Scholar
Irwin, M. R., Olmstead, R., Breen, E. C., Witarama, T., Carrillo, C., Sadeghi, N., … & Cole, S. (2014b). Tai Chi, cellular inflammation, and transcriptome dynamics in breast cancer survivors with insomnia: a randomized controlled trial. Journal of the National Cancer Institute, 50: 295301.Google Scholar
Irwin, M. R., Olmstead, R., Breen, E. C., Witarama, T., Carrillo, C., Sadeghi, N., … & Cole, S. (2015). Cognitive behavioral therapy and Tai Chi reverse cellular and genomic markers of inflammation in late life insomnia: a randomized controlled trial. Biological Psychiatry, 78: 721729.Google Scholar
Irwin, M. R., Olmstead, R., Carrillo, C., Sadeghi, N., Breen, E. C., Witarama, T., … & Nicassio, P. (2014a). Cognitive behavioral therapy vs. Tai Chi for late life insomnia and inflammatory risk: a randomized controlled comparative efficacy trial. Sleep, 37: 15431552.Google Scholar
Irwin, M. R., Olmstead, R., Carrillo, C., Sadeghi, N., Fitzgerald, J. D., Ranganath, V. K., & Nicassio, P. M. (2012). Sleep loss exacerbates fatigue, depression, and pain in rheumatoid arthritis. Sleep, 35: 537543.Google Scholar
Irwin, M. R., Olmstead, R., & Carroll, J. E. (2016). Sleep disturbance, sleep duration, and inflammation: a systematic review and meta-analysis of cohort studies and experimental sleep deprivation. Biological Psychiatry, 80: 4052.Google Scholar
Irwin, M. R., Olmstead, R., & Oxman, M. N. (2007). Augmenting immune responses to varicella zoster virus in older adults: a randomized, controlled trial of Tai Chi. Journal of the American Geriatrics Society, 55: 511517.Google Scholar
Irwin, M. R., Olmstead, R., Valladares, E. M., Breen, E. C., & Ehlers, C. L. (2009). Tumor necrosis factor antagonism normalizes rapid eye movement sleep in alcohol dependence. Biological Psychiatry, 66: 191195.Google Scholar
Irwin, M. R., Thompson, J., Miller, C., Gillin, J. C., & Ziegler, M. (1999). Effects of sleep and sleep deprivation on catecholamine and interleukin-2 levels in humans: clinical implications. Journal of Clinical Endocrinology and Metabolism, 84: 19791985.Google Scholar
Irwin, M. R., Wang, M., Campomayor, C. O., Collado-Hidalgo, A., & Cole, S. (2006). Sleep deprivation and activation of morning levels of cellular and genomic markers of inflammation. Archives of Internal Medicine, 166: 17561762.Google Scholar
Irwin, M. R., Wang, M., Ribeiro, D., Cho, H. J., Olmstead, R., Breen, E. C., … & Cole, S. (2008). Sleep loss activates cellular inflammatory signaling. Biological Psychiatry, 64: 538540.Google Scholar
Jarcho, M. R., Slavich, G. M., Tylova-Stein, H., Wolkowitz, O. M., & Burke, H. M. (2013). Dysregulated diurnal cortisol pattern is associated with glucocorticoid resistance in women with major depressive disorder. Biological Psychology, 93: 150158.Google Scholar
Karin, M. (2006). Nuclear factor-kappaB in cancer development and progression. Nature, 441: 431436.Google Scholar
Kiecolt-Glaser, J. K., Bennett, J. M., Andridge, R., Peng, J., Shapiro, C. L., Malarkey, W. B., … & Glaser, R. (2014). Yoga’s impact on inflammation, mood, and fatigue in breast cancer survivors: a randomized controlled trial. Journal of Clinical Oncology, 32: 10401049.Google Scholar
Kiecolt-Glaser, J. K., Glaser, R., Gravenstein, S., Malarkey, W. B., & Sheridan, J. (1996). Chronic stress alters the immune response to influenza virus vaccine in older adults. Proceedings of the National Academy of Sciences of the USA, 93: 30433047.Google Scholar
Kiecolt-Glaser, J. K., Loving, T. J., Stowell, J. R., Malarkey, W. B., Lemeshow, S., Dickinson, S. L., & Glaser, R. (2005). Hostile marital interactions, proinflammatory cytokine production, and wound healing. Archives of General Psychiatry, 62: 13771384.Google Scholar
Kiecolt-Glaser, J. K., Preacher, K. J., MacCallum, R. C., Atkinson, C., Malarkey, W. B., & Glaser, R. (2003). Chronic stress and age-related increases in the proinflammatory cytokine IL-6. Proceedings of the National Academy of Sciences of the USA, 100: 90909095.Google Scholar
Kubera, M., Symbirtsev, A., Basta-Kaim, A., Borycz, J., Roman, A. Papp, M., & Claesson, M. (1996). Effect of chronic treatment with imipramine on interleukin-1 and interleukin-2 production by splenocytes obtained from rats subjected to a chronic mild stress model of depression. Polish Journal of Pharmacology, 48: 503506.Google Scholar
Lange, T., Dimitrov, S., Bollinger, T., Diekelmann, S., & Born, J. (2011). Sleep after vaccination boosts immunological memory. Journal of Immunology, 187: 283290.Google Scholar
Lange, T., Perras, B., Fehm, H. L., & Born, J. (2003). Sleep enhances the human antibody response to hepatitis A vaccination. Psychosomatic Medicine, 65: 831835.Google Scholar
Lee, H. J., Takemoto, N., Kurata, H., Kamogawa, Y., Miyatake, S., O’Garra, A., & Arai, N. (2000). GATA-3 induces T helper cell type 2 (Th2) cytokine expression and chromatin remodeling in committed Th1 cells. Journal of Experimental Medicine, 192: 105115.Google Scholar
Lotrich, F. E., El-Gabalawi, H., Guenther, L. C., & Ware, C. F. (2011). The role of inflammation in the pathophysiology of depression: different treatments and their effects. Journal of Rheumatology, 88: 4854.Google Scholar
Mangge, H., Kenzian, H., Gallistl, S., Neuwirth, G., Liebmann, P., Kaulfersch, W., … & Schauenstein, K. (1995). Serum cytokines in juvenile rheumatoid arthritis: correlation with conventional inflammation parameters and clinical subtypes. Arthritis and Rheumatism, 38: 211220.Google Scholar
Mantovani, A., Allavena, P., Sica, A., & Balkwill, F. (2008). Cancer-related inflammation. Nature, 454: 436444.Google Scholar
Matthews, K. A., Schott, L. L., Bromberger, J. T., Cyranowski, J. M., Everson-Rose, S. A., & Sowers, M. (2010). Are there bi-directional associations between depressive symptoms and C-reactive protein in mid-life women? Brain, Behavior, and Immunity, 24: 96101.Google Scholar
Matzinger, P. (2007). Friendly and dangerous signals: is the tissue in control? Nature Immunology, 8: 11–3.Google Scholar
Mays, J. W., Bailey, M. T., Hunzeker, J. T., Powell, N. D., Papenfuss, T., Karlsson, E. A., … & Sheridan, J. F. (2010). Influenza virus-specific immunological memory is enhanced by repeated social defeat. Journal of Immunology, 184: 20142025.Google Scholar
McEwen, B. S. (2007). Physiology and neurobiology of stress and adaptation: central role of the brain. Physiological Reviews, 87: 873904.Google Scholar
Medzhitov, R. (2008). Origin and physiological roles of inflammation. Nature, 454: 428435.Google Scholar
Meier-Ewert, H. K., Ridker, P. M., Rifai, N., Regan, M. M., Price, M. J., Dinges, D. F., & Mullington, J. M. (2004). Effect of sleep loss on C-reactive protein, an inflammatory marker of cardiovascular risk. Journal of the American College of Cardiology, 43: 678683.Google Scholar
Miller, A. H., Ancoli-Israel, S., Bower, J. E., Capuron, L., & Irwin, M. R. (2008a). Neuroendocrine-immune mechanisms of behavioral comorbidities in patients with cancer. Journal of Clinical Oncology, 26: 971982.Google Scholar
Miller, A. H., Maletic, V., & Raison, C. L. (2009a). Inflammation and its discontents: the role of cytokines in the pathophysiology of major depression. Biological Psychiatry, 65: 732741.Google Scholar
Miller, G. E. & Chen, E. (2010). Harsh family climate in early life presages the emergence of a proinflammatory phenotype in adolescence. Psychological Science, 21: 848856.Google Scholar
Miller, G. E., Chen, E., Fok, A. K., Walker, H., Lim, A., Nicholls, E. F., … & Kobor, M. S. (2009b). Low early-life social class leaves a biological residue manifested by decreased glucocorticoid and increased proinflammatory signaling. Proceedings of the National Academy of Sciences of the USA, 106: 1471614721.Google Scholar
Miller, G. E., Chen, E., Sze, J., Marin, T., Arevalo, J. M., Doll, R., … & Cole, S. W. (2008b). A functional genomic fingerprint of chronic stress in humans: blunted glucocorticoid and increased NF-kappaB signaling. Biological Psychiatry, 64: 266272.Google Scholar
Miller, G. E. & Cohen, S. (2001). Psychological interventions and the immune system: a meta-analytic review and critique. Health Psychology, 20: 4763.Google Scholar
Miller, G. E., Cohen, S., Pressman, S. D., Barkin, A., Rabin, B., & Treanor, J. J. (2004). Psychological stress and antibody response to influenza vaccination: when is the critical period for stress, and how does it get inside the body? Psychosomatic Medicine, 66: 215223.Google Scholar
Miller, G. E. & Cole, S. W. (1998). Social relationships and the progression of human immunodeficiency virus infection: a review of evidence and possible underlying mechanisms. Annals of Behavioral Medicine, 20: 181189.Google Scholar
Miller, N. E. (1964). Some psychophysiological studies of the motivation and of the behavioral effects of illness. Bulletin of the British Psychological Society, 17: 120.Google Scholar
Moons, W. G., Eisenberger, N. I., & Taylor, S. E. (2010). Anger and fear responses to stress have different biological profiles. Brain, Behavior, and Immunity, 24: 215219.Google Scholar
Morgan, N., Irwin, M. R., Chung, M., & Wang, C. (2014). The effects of mind–body therapies on the immune system: meta-analysis. PLoS One, 9: e100903.Google Scholar
Motivala, S. & Irwin, M.R. (2007). Sleep and immunity: cytokine pathways linking sleep and health outcomes. Current Directions in Psychological Science, 16: 2125.Google Scholar
Mullington, J., Korth, C., Hermann, D. M., Orth, A., Galanos, C., Holsboer, F., & Pollmächer, T. (2000). Dose-dependent effects of endotoxin on human sleep. American Journal of Physiology. Regulatory, Integrative and Comparative Physiology, 278: R947R955.Google Scholar
Murphy, K. (2001). Janeway’s Immunobiology. New York: Garland.Google Scholar
Murphy, M. L. M., Slavich, G. M., Chen, E., & Miller, G. E. (2015). Targeted rejection predicts decreased anti-inflammatory gene expression and increased symptom severity in youth with asthma. Psychological Science, 26: 111121.Google Scholar
Murphy, M. L. M., Slavich, G. M., Rohleder, N., & Miller, G. E. (2013). Targeted rejection triggers differential pro- and anti-inflammatory gene expression in adolescents as a function of social status. Clinical Psychological Science, 1: 3040.Google Scholar
Murray, D. R., Irwin, M., Rearden, C. A., Ziegler, M., Motulsky, H., & Maisel, M. S. (1992). Sympathetic and immune interactions during dynamic exercise: mediation via Beta2-adrenergic dependent mechanism. Circulation, 83: 203213.Google Scholar
Nance, D. M. & Sanders, V. M. (2007). Autonomic innervation and regulation of the immune system (1987–2007). Brain, Behavior, and Immunity, 21: 736745.Google Scholar
Nicklas, B. J., Hsu, F. C., Brinkley, T. J., Church, T., Goodpaster, B. H., Kritchevsky, S. B., & Pahor, M. (2008). Exercise training and plasma C-reactive protein and interleukin-6 in elderly people. Journal of the American Geriatrics Society, 56: 20452052.Google Scholar
O’Connor, M. F., Bower, J. E., Cho, H. J., Creswell, J. D., Dimitrov, S., Hamby, M. E., … & Irwin, M. R. (2009). To assess, to control, to exclude: effects of biobehavioral factors on circulating inflammatory markers. Brain, Behavior, and Immunity, 23: 887897.Google Scholar
O’Connor, M. F., Motivala, S., Valladares, E. M., Olmstead, R. E., & Irwin, M. R. (2007). Sex differences in monocyte expression of IL-6: role of autonomic mechanisms. American Journal of Physiology. Regulatory, Integrative and Comparative Physiology, 293: R145R151.Google Scholar
O’Donovan, A., Slavich, G. M., Epel, E. S., & Neylen, T. C. (2013). Exaggerated neurobiological sensitivity to threat as a mechanism linking anxiety with increased risk for diseases of aging. Neuroscience & Biobehavioral Reviews, 37: 96108.Google Scholar
Pace, T. W., Hu, F., & Miller, A. H. (2007). Cytokine-effects on glucocorticoid receptor function: relevance to glucocorticoid resistance and the pathophysiology and treatment of major depression. Brain, Behavior, and Immunity, 21: 919.Google Scholar
Pace, T. W., Mletzko, T. C., Alagbe, O., Musselman, D. L., Nemeroff, C. B., Miller, A. H., & Heim, C. M. (2006). Increased stress-induced inflammatory responses in male patients with major depression and increased early life stress. American Journal of Psychiatry, 163: 16301633.Google Scholar
Panina-Bordignon, P., Mazzeo, D., Di Lucia, P., D’Ambrosio, D., Lang, R., Fabbri, L., … & Sinigaglia, F. (1997). β2-agonists prevent Th1 development by selective inhibition of interleukin 12. Journal of Clinical Investigation, 100: 15131519.Google Scholar
Patel, S. R., Malhotra, A., Gao, X., Hu, F. B., Neuman, M. I., & Fawzi, W. W. (2012). A prospective study of sleep duration and pneumonia risk in women. Sleep, 35: 97101.Google Scholar
Patten, S. B., Williams, J. V. A., Lavorato, D. H., Modgill, G., Jetté, N., & Eliasziw, M. (2008). Major depression as a risk factor for chronic disease incidence: longitudinal analyses in a general population cohort. General Hospital Psychiatry, 30: 407413.Google Scholar
Phillips, A. C., Carroll, D., Burns, V. E., & Drayson, M. (2009). Cardiovascular activity and the antibody response to vaccination. Journal of Psychosomatic Research, 67: 3743.Google Scholar
Polsky, D., Doshi, J. A., Marcus, S., Oslin, D., Rothbard, A., Thomas, N., & Thompson, C. L. (2005). Long-term risk for depressive symptoms after a medical diagnosis. Archives of Internal Medicine, 165: 12601266.Google Scholar
Powell, N. D., Mays, J. W., Bailey, M. T., Hanke, M. L., & Sheridan, J. F. (2011). Immunogenic dendritic cells primed by social defeat enhance adaptive immunity to influenza A virus. Brain, Behavior, and Immunity, 25 (1): 4652.Google Scholar
Prather, A. A., Hall, M., Fury, J. M., Ross, D. C., Muldoon, M. F., Cohen, S., & Marsland, A. L. (2012). Sleep and antibody response to hepatitis B vaccination. Sleep, 35: 10631069.Google Scholar
Raetz, C. R. & Whitfield, C. (2002). Lipopolysaccharide endotoxins. Annual Review of Biochemistry, 71: 635700.Google Scholar
Raison, C. L. & Miller, A. H. (2003). Depression in cancer: new developments regarding diagnosis and treatment. Biological Psychiatry, 54: 283294.Google Scholar
Raison, C. L., Rutherford, R. E., Woolwine, B. J., Shuo, C., Schettler, P., Drake, D. F., … & Miller, A. H. (2013). A randomized controlled trial of the tumor necrosis factor antagonist infliximab for treatment-resistant depression: the role of baseline inflammatory biomarkers. JAMA Psychiatry, 70: 3141.Google Scholar
Raison, C. L., Rye, D. B., Woolwine, B. J., Vogt, G. J., Bautista, B. M., Spivey, J. R., & Miller, A. H. (2010). Chronic interferon-alpha administration disrupts sleep continuity and depth in patients with hepatitis C: association with fatigue, motor slowing, and increased evening cortisol. Biological Psychiatry, 68: 942949.Google Scholar
Renegar, K. B., Floyd, R. A., & Krueger, J. M. (1998). Effects of short-term sleep deprivation on murine immunity to influenza virus in young adult and senescent mice. Sleep, 21: 241248.Google Scholar
Rhen, T. & Cidlowski, J. A. (2005). Antiinflammatory action of glucocorticoids: new mechanisms for old drugs. New England Journal of Medicine, 353: 17111723.Google Scholar
Richlin, V. A., Arevalo, J. M., Zack, J. A., & Cole, S. W. (2004). Stress-induced enhancement of NF-kappaB DNA-binding in the peripheral blood leukocyte pool: effects of lymphocyte redistribution. Brain, Behavior, and Immunity, 18: 231237.Google Scholar
Rooney, A. G., McNamara, S., Mackinnon, M., Fraser, M., Rampling, R., Carson, A., & Grant, R. (2011). Frequency, clinical associations, and longitudinal course of major depressive disorder in adults with cerebral glioma. Journal of Clinical Oncology, 29: 43074312.Google Scholar
Sapolsky, R., Rivier, C., Yamamoto, G., Plotsky, P., & Vale, W. (1987). Interleukin-1 stimulates the secretion of hypothalamic corticotropin-releasing factor. Science, 238: 522524.Google Scholar
Schleifer, S. J., Keller, S. E., & Bartlett, J. A. (1999). Depression and immunity: clinical factors and therapeutic course. Psychiatry Research, 85: 6369.Google Scholar
Schultze-Florey, C. R., Martínez-Maza, O., Magpantay, L., Breen, E. C., Irwin, M. R., Gündel, H., & O’Connor, M.-F. (2012). When grief makes you sick: bereavement induced systemic inflammation is a question of genotype. Brain, Behavior, and Immunity, 26: 10661071.Google Scholar
Slavich, G. M. (2015). Understanding inflammation, its regulation, and relevance for health: a top scientific and public priority. Brain, Behavior, and Immunity, 45: 1314.Google Scholar
Slavich, G. M. & Cole, S. W. (2013). The emerging field of human social genomics. Clinical Psychological Science, 1: 331348.Google Scholar
Slavich, G. M. & Irwin, M. R. (2014). From stress to inflammation and major depressive disorder: a social signal transduction theory of depression. Psychological Bulletin, 140: 774815.Google Scholar
Slavich, G. M., Monroe, S. M., & Gotlib, I. H. (2011). Early parental loss and depression history: associations with recent life stress in major depressive disorder. Journal of Psychiatric Research, 45: 11461152.Google Scholar
Slavich, G. M., O’Donovan, A., Epel, E. S., & Kemeny, M. E. (2010). Black sheep get the blues: a psychobiological model of social rejection and depression. Neuroscience & Biobehavioral Reviews, 35: 3945.Google Scholar
Slavich, G. M., Tartter, M. A., Brennan, P. A., & Hammen, C. (2014). Endogenous opioid system influences depressive reactions to socially painful targeted rejection life events. Psychoneuroendocrinology, 49: 141149.Google Scholar
Slavich, G. M., Thornton, T., Torres, L. D., Monroe, S. M., & Gotlib, I. H. (2009). Targeted rejection predicts hastened onset of major depression. Journal of Social and Clinical Psychology, 28: 223243.Google Scholar
Sloan, E. K., Capitanio, J. P., Tarara, R. P., Mendoza, S. P., Mason, W. A., & Cole, S. W. (2007). Social stress enhances sympathetic innervation of primate lymph nodes: mechanisms and implications for viral pathogenesis. Journal of Neuroscience, 27: 88578865.Google Scholar
Smith, C. W., Marlin, S. D., Rothlein, R., Toman, C., & Anderson, D. C. (1989). Cooperative interactions of LFA-1 and Mac-1 with intercellular adhesion molecule-1 in facilitating adherence and transendothelial migration of human neutrophils in vitro. Journal of Clinical Investigation, 83: 20082017.Google Scholar
Spiegel, K., Sheridan, J. F., & Van Cauter, E. (2002). Effect of sleep deprivation on response to immunization. Journal of the American Medical Association, 288: 14711472.Google Scholar
Thomas, K. S., Motivala, S., Olmstead, R., & Irwin, M. R. (2011). Sleep depth and fatigue: role of cellular inflammatory activation. Brain, Behavior, and Immunity, 25: 5358.Google Scholar
Torres, M. A., Pace, T. W., Liu, T., Felger, J. C., Mister, D., Doho, G. H., … & Miller, A. H. (2013). Predictors of depression in breast cancer patients treated with radiation: role of prior chemotherapy and nuclear factor kappa B. Cancer, 119: 19511959.Google Scholar
Tyring, S., Gottlieb, A., Papp, K., Gordon, K., Leonardi, C., Wang, A., … & Krishnan, R. (2006). Etanercept and clinical outcomes, fatigue, and depression in psoriasis: double-blind placebo-controlled randomised phase III trial. Lancet, 367: 2935.Google Scholar
Vedhara, K., Cox, N. K. M., Wilcock, G. K., Perks, P., Hunt, M., Anderson, S., … & Shanks, N. M. (1999). Chronic stress in elderly carers of dementia patients and antibody response to influenza vaccination. Lancet, 353: 627631.Google Scholar
Wang, C., Banuru, R., Ramel, J., Kupelnick, B., Scott, T., & Schmid, C. H. (2010). Tai Chi on psychological well-being: systematic review and meta-analysis. BMC Complementary and Alternative Medicine, 10: 23.Google Scholar
Wang, C., Collet, J. P., & Lau, J. (2004). The effect of Tai Chi on health outcomes in patients with chronic conditions: a systematic review. Archives of Internal Medicine, 164: 493501.Google Scholar
Watkins, L. R. & Maier, S. F. (1999). Implications of immune-to-brain communication for sickness and pain. Proceedings of the National Academy of Sciences of the USA, 96: 77107713.Google Scholar
Wohleb, E. S., Hanke, M. L., Corona, A. W., Powell, N. D., Stiner, L. T. M., Bailey, M. T., … & Sheridan, J. F. (2011). Beta-adrenergic receptor antagonism prevents anxiety-like behavior and microglial reactivity induced by repeated social defeat. Journal of Neuroscience, 31: 62776288.Google Scholar
Zautra, A. J., Davis, M. C., Reich, J. W., Nicassario, P., Tennen, H., Finan, P., … & Irwin, M. R. (2008). Comparison of cognitive behavioral and mindfulness meditation interventions on adaptation to rheumatoid arthritis for patients with and without history of recurrent depression. Journal of Consulting and Clinical Psychology, 76: 408421.Google Scholar
Zautra, A. J., Yocum, D. C., Villanueva, I., Smith, B., Davis, M. C., Attrep, J., & Irwin, M. (2004). Immune activation and depression in women with rheumatoid arthritis. Journal of Rheumatology, 31: 457463.Google Scholar

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
×