Skip to main content

Principles of Electrodermal Phenomena

  • Chapter
  • First Online:
Electrodermal Activity

Abstract

Since the 1880s, when psychological factors related to electrodermal phenomena were first observed, electrodermal recording has become one of the most frequently used biosignals in psychophysiology. The major reason for its popularity is the ease of obtaining a distinct electrodermal response (EDR), the intensity of which seems apparently related to stimulus intensity and/or its psychological significance. Electrodermal recording is possible with rather inexpensive equipment, not only in the laboratory but also under less controlled field conditions.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 189.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 249.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 249.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    Dermal stems from Latin: derma = true skin, see Table 1.2, Sect. 1.2.1.1.

  2. 2.

    Abbreviations are determined by the first letter of the words: skin, potential, resistance, and conductance. Unfortunately, the commission overlooked that the abbreviation C is already reserved in physics for capacitance, and G is used for conductance, according to SI units (Sect. 1.4.1.1). The terminology used in AC methodology is somewhat more complicated. Edelberg (1972a) proposed A for admittance, and Z – the last letter in the alphabet – for the reciprocal unit, impedance. While the latter abbreviation was kept, admittance is abbreviated today by Y, the penultimate letter of the alphabet.

  3. 3.

    Note that the dermis is also known as the corium (Latin for tauter skin) and that sometimes the term cutis is used only for the dermal part of skin (cutis vera, see Table 1.2).

  4. 4.

    For example, Birgersson et al. (2011) performed a literature review on the skin thickness at the female volar forearm, revealing a stratum corneum thickness of 14 ± 3 μm compared to 1.2 ± .2 mm for the subcutis.

  5. 5.

    Some authors (e.g., Orfanos, 1972) proposed a division into three layers, which may have been dependent upon different microscopic technology.

  6. 6.

    This transitional layer, with an overall thickness of around 1 μm, is unusually thin in comparison with the underlying Malpighian layer and the overlying part of the epidermis. Orfanos (1972) has hence suggested that this layer should not be subdivided, but should be named on the whole as the stratum intermedium (i.e., intermediate) instead. However, sometimes only the granular layer is distinguished as the transitional zone (Jarrett, 1973a).

  7. 7.

    The process of keratinization begins during mitosis via forming of the so-called tonofilaments, which are thin intraplasmatic fibers. In the upper layers, they are transformed into bundles of tonofibrillas having a greater density. With the keratohyaline generated in the lower stratum intermedium, these fibrillas merge to form complexes, which are converted in the upper stratum intermedium from one cell layer to another into epidermal keratin through changes in the cellular milieu. The tonofilaments are approximately parallel to the surfaces of the flattened cells, but are not in contact with the desmosomes, which are the intercellular contact zones of the keratinocytes. In the prickle cell layer, the number of desmosomes decreases and the intercellular spaces enlarge. This enables other cells, e.g., the melanocytes, to change their positions through movements.

  8. 8.

    Sato et al. (1989) give slightly different numbers for the eccrine sweat glands: 1.6–4.0 million glands in total, with densities per cm2 of 64 on the back, 108 on the forearm, 181 on the forehead, and 600–700 on the palms and soles. For age, gender, and ethnic differences, see Sect. 2.4.3.

  9. 9.

    Some authors (e.g., Thiele, 1981a) use the term “atrichial” (from Greek: τρίχα = hair) for the eccrine glands, referring to these glands being not associated with hair follicles, which most apocrine glands are (Venables & Christie, 1973, p. 19).

  10. 10.

    Sweat glands are found in the footpads of many species (Edelberg, 1972a, p. 378). Wang (1964) presumes that the cat’s sweat glands are apocrine instead of eccrine, but this view remains controversial (Edelberg, 1972a). Sato (1977) found it unlikely that the cat’s or the rat’s sweat glands had the capability to reabsorb ductal NaCl (Sect. 1.3.3.1), while those from the monkey’s paw closely resemble human eccrine sweat glands. Thus, generalization from primate results to human species is more tenable than from cats or rats (Roy, Sequeira, & Delerm, 1993). Types and distributions of sweat glands in different species are summed up in a description by Fowles (1986a, p. 54); more details are given by Weiner and Hellmann (1960).

  11. 11.

    Only few human apocrine glands possess an ANS innervation (e.g., those in the axilla); they may be partly or even mainly under the control of circulating adrenaline (Weiner & Hellmann, 1960).

  12. 12.

    It is still unclear whether specific vasodilative fibers exist in the human skin (which are known from cats), or if vasodilatation is secondary to sweating (Wallin, 1992).

  13. 13.

    The search for origins of the skin’s vegetative activity is further complicated by the fact that transmitters such as noradrenaline, which acts vasoconstrictory on the peripheral blood vessels (Sect. 1.2.1.3), are also circulating freely in the blood.

  14. 14.

    Sometimes also called sudorisecretory (sweat secretion eliciting) fibers.

  15. 15.

    Sato and Sato (1981) used an in vitro preparation of single monkey palm eccrine sweat glands to demonstrate its reactivity to both cholinergic and andrenergic agents. They found that the maximum sweat rate was highest after cholinergic stimulation.

  16. 16.

    Recording of action potentials from nerves with tungsten microelectrodes having tip diameters of a few μm, which are inserted through the intact skin into an underlying nerve, with a reference electrode placed subcutaneously in 1–2 cm distance (Wallin, 1992).

  17. 17.

    Microneurography from the tibial nerve has been used by Nishiyama et al. (2001), together with videomicroscopy for recording sweat secretion of individual glands on the sole of four male individuals. Only 46% of the suprathreshold sudomotor nerve bursts elicited sweat secretion, but the number of sweat glands recruited was linearly related to the amplitude of sudomotor bursts. Hence, sweat secretion is primarily dependent on the intensity of sudomotor neural activity; however, the microenvironment of sweat glands may also play a role.

  18. 18.

    Nishiyama et al. (2001), in their study described in the previous footnote, also recorded SPRs, finding that sudomotor bursts were followed by SPRs with latencies of 1.33 ± .33 s. The latency of sweat secretion was 2.29 ± .03 s, and the sweat expulsion latency was 3.22 ± .03 s.

  19. 19.

    Recently, the existence of the limbic system has been much debated (e.g., LeDoux, 1996). However, since most of the literature on CNS elicitation of electrodermal phenomena used this expression (see Sect 1.3.4.1 “Subcortical Control of EDA”), it will be used in the present book as well.

  20. 20.

    An appropriate double innervation may exist in the region of the mouth floor which corresponds to dermatome C3 (Table 1.3, Sect. 1.3.2.1).

  21. 21.

    It is also important for cardiac functions related to blood pressure. The metabolic processes that take place during sweat secretion, including the possible role of Ca as a second messenger in cholinergic sweating, are discussed in detail by Sato (1977) and summed up by Fowles (1986a).

  22. 22.

    Sato (1977) generally questions the hypothesis of expulsion of already existing sweat by adrenergic stimulation. In his in vitro studies with isolated monkey sweat glands, he found so little preformed sweat in the lumen that an initial myoepithelial contraction could not expel an appreciable amount of sweat. Furthermore, Nicolaidis and Sivadjian (1972) used forehead sites for recording, so their observations may not be generalized to palmar sweat secretion.

  23. 23.

    In deviation from the classical dermatological hypothesis, Tregear (1966) presumed that the sweat gland ducts, as well as the hair follicles, were of no importance for the amount of the body’s water loss. He referred to observations that the palmar skin, having three times the density of sweat glands compared to the rest of the body, is not very permeable to fluids, and that individuals without any sweat glands show as much insensible perspiration in a cool environment as normal individuals.

  24. 24.

    The specific meaning of these different kinds of sweating for EDA is as yet unexplored, except those of the so-called emotional-sweating type. They are reported here for the sake of completion and with respect to possible future electrodermal research.

  25. 25.

    One major difference between sweating in cats and humans may be that there is probably no thermal sweating in the cat (Jänig et al., 1983). Therefore, Roy, Sequeira, and Delerm (1993), hypothesized that the sweating of the cat’s footpad could be analogous to “emotional sweating” in humans (Sects. 1.3.2.4 and 1.3.3.3).

  26. 26.

    Skin conductance was recorded bilaterally with 0.5 V constant voltage from the medial phalanges, using Ag/AgCl electrodes filled with 0.5% KCl in an agar medium (Sect. 2.2.2.5). A series of six 75 dB tones (1.311 Hz, 1 s, ISI randomized between 35 and 50 s) and four reorienting stimuli for probing OR reinstatement after habituation (Sect. 3.1.1) were applied. Left and right hand SC-OR was significantly related to left and right prefrontal areas (r = 0.44–0.60), area of the pons (r = 0.43–0.54) and left but not right temporal/amygdalar area (r = 0.47–0.53). No significant correlations were found with the area of the cerebellum, nonfrontal cortical areas, and the medial prefrontal cortex.

  27. 27.

    Animal research on hypothalamo-limbic and cortical control of EDA has been summarized by Sequeira and Roy (1993) in their Table 1.1, pp. 97–99.

  28. 28.

    In general, the question of ipsi- or contralaterality of inhibitory vs. excitatory influences on electrodermal phenomena remains unresolved (Sect. 3.1.4.2).

  29. 29.

    The bilateral medial prefrontal cortex is subsumed by the VMPFC, an area that is most consistently associated with missing SCR in patients with appropriate lesions. The possible role of brain areas for EDA biofeedback is discussed in Sect. 3.1.2.3.

  30. 30.

    In the studies of this group, patients were diagnosed by MRI or X-ray computerized tomography, and SC was recorded with standard methodology (Sect. 2.2.7) from thenar/hypothenar sites.

  31. 31.

    Recorded with standard methodology from 36 brain-damaged (15 females, 21 males) compared to 20 matched normal participants (7 females, 13 males).

  32. 32.

    Measured with a biofeedback system via silver electrodes and KCl electrolyte cream from the palmar surface of the left hand. Their method of recording SC within an fMRI was described by Critchley et al. (2000), see Sect. 2.2.3.5.

  33. 33.

    However, Langworthy and Richter (1930) could elicit EDRs and other autonomic responses by stimulating the pyramidal tract in cats. Roy, Sequeira-Martinho, and Brochard (1984) suggested this was due to the collaterals from the pyramidal tract reaching the RF, by which reticular elicitation of EDA has been mediated.

  34. 34.

    Sequeira and Roy (1993) managed to elicit EDRs in cats by stimulating both motor and premotor cortical areas electrically, even in a “pyramidal preparation,” where all descending pathways except the pyramidal tracts were interrupted at bulbar level.

  35. 35.

    White noise and videos of snakes with or without electric shocks delivered to the right hand. Laterality could not be evaluated since EDA was only measured on the left hand. SC was recorded with standard methodology (Sect. 2.2.7); the amplitude criterion was set to 0.05 μS.

  36. 36.

    Wilcott (1963) showed that nonpalmar areas of skin that are normally regarded as thermoregulatory also took part in emotional sweating during a stressful situation (Sect. 3.2.2.2).

  37. 37.

    More recently, Lademann, Jacobi, Surber, Weighman, and Fluhr (2008) showed that up to around 80 strippings may be required to fully remove the stratum corneum.

  38. 38.

    From their lesion and stimulation research with cats, Sequeira and Roy (1993) concluded that a direct corticospinal sudomotor pathway may exist, contributing to the corticospinal control of final autonomic adjustments, in particular during grasping.

  39. 39.

    In Anglo-American papers, there was a widespread custom for using the unit “mho,” i.e., the mirror image of “ohm,” instead. Meanwhile, S was introduced as the SI unit for conductance. Despite Venables and Christie (1980) argued for the continued usage of mho, in the last 20–30 years most researchers have used the unit S for conductance. The unit mho corresponds 1:1 to the unit μS.

  40. 40.

    During the following considerations, the inner resistance of the voltage source should always be negligible for the reason of simplification.

  41. 41.

    These connections can also be elucidated through a depiction with complex numbers. There R(f) is taken as the real part and X(f) as the imaginary part of a complex function. This depiction, which is preferred in electrophysics and in systems theory, is shown in the lower part of Fig. 1.10.

  42. 42.

    Edelberg (1971) reported microelectrode recordings which provide evidence for the existence of an electrical barrier layer in the deeper layers of the epidermis. The SRL measured via a microelectrode, which had been slowly pushed into the epidermis, showed a slow continuous decrease at the beginning. If a certain point had been passed at which the participant first reported weak pain, SRL suddenly decreased until only the electrode resistance itself was present. The depth of the appropriate layer is 350 μm at the palm and 50 μm at the forearm.

  43. 43.

    Edelberg (1971) first suggested a second barrier membrane at the dermal/epidermal boundary. According to Fowles (1974), he later preferred the ductal wall at the height of the stratum germinativum as the locus of this second membrane.

  44. 44.

    The SPRs in cats differ from those seen in humans in that they show only a monophasic negative SPR (Edelberg, 1973a), which reaches its peak amplitude very quickly (for differences between species, see also Footnote 10, Sect. 1.2.3).

  45. 45.

    Taking into consideration that R 1 = constant and with application of the quotient rule for differentiation. The corresponding conductance equation is given in Boucsein, Baltissen, and Euler (1984).

  46. 46.

    As shown in the left-hand panel of Fig. 1.15, Montagu and Coles (1966) in principle acknowledged the necessity of including capacitive elements into electrodermal modeling.

  47. 47.

    However, since all the potential sources are “leaky” capacitors (see comments on Fig. 1.17 below), the parallel resistive component can conduct DC even when the capacitive component is fully charged.

  48. 48.

    For the sake of simplification, Fowles omitted a fourth pathway in which the current flows into the duct from the corneum and then along the two sweat gland pathways.

  49. 49.

    Observed in 10 healthy male participants who were subjected to handclaps and bursts of 85 dB white noise. Palmar sweating was recorded with a capacitance hygrometry technique, using a capsule mounted on the left hypothenar eminence close to the active SP electrode. The capsule was ventilated with dry nitrogen. For SP recording, Ag/AgCl electrodes filled with isotonic electrode cream made from Unibase were used, the neutral electrode being attached to a lightly abraded ventral site of the left forearm.

  50. 50.

    Recently, Birgersson et al. (2011) refined the AC-recording technique as used by the Yamamoto group by means of electrical impedance spectroscopy. They also proposed a mathematical three-layer model of the skin, comprising (1) the stratum corneum, (2) the viable skin, which includes the other epidermal layers and the dermis, and (3) the fatty tissue of the hypodermis (Sect. 1.2.1.2). A noninvasive gold-plated probe, featuring two voltage injection electrodes, one current detector and a circular guard electrode (for diminishing the impact of surface leaking currents) was designed for carrying out two-point measurements with sinusoidal AC of varying frequencies. Recordings were taking from the volar forearm of a homogenous group of 60 young females. The skin site was soaked with a physiological saline solution (0.9% NaCl) for 1 min before the first recording was performed. Subsequent recordings were carried out at 5 min intervals up until 30 min in an ambient temperature of 21.7°C (±3°C) and a relative humidity of 36% (±7%). Impedance magnitudes (in kΩ) and the phase angle ϕ (in degrees) were measured at five different current penetration depths (by varying the voltage at the second current injection electrode from 5 to 50 mV) at 35 frequencies logarithmically distributed between 1.0 kHz and 2.5 MHz. In addition, the impact of NaCl concentration (0.9–18% NaCl) and soaking time was investigated in ten participants. Preliminary results revealed that – in the frequency range – the amount of current passing through the hypodermis (subcutis) can be regarded as negligible. As to be expected, the measured SZ varied considerably with the saline concentration and soaking time.

References

  • Adams, T. (1966). Characteristics of eccrine sweat gland activity in the footpad of the cat. Journal of Applied Physiology, 21, 1004–1012.

    PubMed  Google Scholar 

  • Adams, T., & Hunter, W. S. (1969). Modification of skin mechanical properties by eccrine sweat gland activity. Journal of Applied Physiology, 26, 417–419.

    PubMed  Google Scholar 

  • Allen, J. A., Armstrong, J. E., & Roddie, I. C. (1973). The regional distribution of emotional sweating in man. The Journal of Physiology, 235, 749–759.

    PubMed  Google Scholar 

  • Arthur, R. P., & Shelley, W. B. (1959). The innervation of human epidermis. The Journal of Investigative Dermatology, 32, 397–411.

    PubMed  Google Scholar 

  • Bagshaw, M. H., Kimble, D. P., & Pribram, K. H. (1965). The GSR of monkeys during orienting and habituation and after ablation of the amygdala, hippocampus and inferotemporal cortex. Neuropsychologia, 3, 111–119.

    Google Scholar 

  • Ba-M’hamed-Bennis, S., Sequeira- Martinho, H., Freixa i Baqué, E., & Roy, J.-C. (1985). Skin potential responses elicited by reticular stimulation are not lateralized in the cat. Biological Psychology, 21, 250–251.

    Google Scholar 

  • Bard, P. (1960). Anatomical organization of the central nervous system in relation to control of the heart and blood vessels. Physiological Reviews, 4(Suppl. 4), 3–26.

    PubMed  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. The Journal of Neuroscience, 19, 5473–5481.

    PubMed  Google Scholar 

  • Bechara, A., Tranel, D., Damasio, H., & Damasio, A. R. (1996). Failure to respond automatically to anticipated future outcomes following damage to prefrontal cortex. Cerebral Cortex, 6, 215–225.

    PubMed  Google Scholar 

  • Benedek, M., & Kaernbach, C. (2010). Decomposition of skin conductance data by means of nonnegative deconvolution. Psychophysiology, 47, 647–658.

    PubMed  Google Scholar 

  • Bijman, J. (1987). Transport processes in the eccrine sweat gland. Kidney International. Supplement, 21, S109–S112.

    PubMed  Google Scholar 

  • Bing, H. I., & Skouby, A. P. (1950). Sensitization of cold receptors by substances with acetylcholine effect. Acta Physiologica Scandinavica, 21, 286–302.

    PubMed  Google Scholar 

  • Birgersson, U., Birgersson, E., Aberg, P., Nicander, I., & Ollmar, S. (2011). Non-invasive bioimpedance of intact skin: Mathematical modeling and experiments. Physiological Measurement, 32, 1–18.

    PubMed  Google Scholar 

  • Bloch, V. (1952). Nouveaux aspects de la méthode psychogalvanique ou électrodermographique (EDG) comme critère des tensions affectives. L’Année Psychologique, 52, 329–362.

    PubMed  Google Scholar 

  • Bloch, V. (1965). Le contrôle central de l’activité électrodermale. Journal de Physiologie, 57, 1–132.

    PubMed  Google Scholar 

  • Boucsein, W. (1988). Elektrodermale Aktivität. Grundlagen, Methoden und Anwendungen. Berlin: Springer.

    Google Scholar 

  • Boucsein, W. (1992). Electrodermal Activity. New York: Plenum.

    Google Scholar 

  • Boucsein, W., & Backs, R. W. (2000). Engineering psychophysiology as a discipline: Historical and theoretical aspects. In R. W. Backs & W. Boucsein (Eds.), Engineering psychophysiology. Issues and applications (pp. 3–30). Mahwah, NJ: Lawrence Erlbaum.

    Google Scholar 

  • Boucsein, W., Baltissen, R., & Euler, W. (1984). Dependence of skin conductance reactions and skin resistance reactions on previous level. Psychophysiology, 21, 212–218.

    PubMed  Google Scholar 

  • Boucsein, W., Schaefer, F., & Neijenhuisen, H. (1989). Continuous recording of impedance and phase angle during electrodermal reactions and the locus of impedance change. Psychophysiology, 26, 369–376.

    PubMed  Google Scholar 

  • Braus, H., & Elze, C. (1960). Anatomie des Menschen (Human Anatomy) (Vol. 3). Berlin: Springer.

    Google Scholar 

  • Brown, C. C. (1967). A proposed standard nomenclature for psychophysiological measures. Psychophysiology, 4, 260–264.

    PubMed  Google Scholar 

  • Brown, B. H., Bygrave, C., Robinson, P., & Henderson, H. P. (1980). A critique of the use of a thermal clearance probe for the measurement of skin blood flow. Clinical Physics and Physiological Measurement, 1, 237.

    Google Scholar 

  • Bundy, R. S., & Fitzgerald, H. E. (1975). Stimulus specificity of electrodermal recovery time: An examination and reinterpretation of the evidence. Psychophysiology, 12, 406–411.

    PubMed  Google Scholar 

  • Burbank, D. P., & Webster, J. G. (1978). Reducing skin potential motion artifact by skin abrasion. Medical & Biological Engineering & Computing, 16, 31–38.

    Google Scholar 

  • Campbell, S. D., Kraning, K. K., Schibli, E. G., & Momii, S. T. (1977). Hydration characteristics and electrical resistivity of stratum corneum using a noninvasive four-point microelectrode method. The Journal of Investigative Dermatology, 69, 290–295.

    PubMed  Google Scholar 

  • Conklin, J. E. (1951). Three factors affecting the general level of electrical skin-resistance. The American Journal of Psychology, 64, 78–86.

    PubMed  Google Scholar 

  • Critchley, H. D. (2002). Electrodermal responses: What happens in the brain. The Neuroscientist, 8, 132–142.

    PubMed  Google Scholar 

  • Critchley, H. D., Elliott, 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. The Journal of Neuroscience, 20, 3033–3040.

    PubMed  Google Scholar 

  • Culp, W. C., & Edelberg, R. (1966). Regional response specifity in the electrodermal reflex. Perceptual and Motor Skills, 23, 623–627.

    Google Scholar 

  • Darrow, C. W. (1933). The functional significance of the galvanic skin reflex and perspiration on the backs and palms of the hands. Psychological Bulletin, 30, 712.

    Google Scholar 

  • Darrow, C. W. (1937a). Neural mechanisms controlling the palmar galvanic skin reflex and palmar sweating. Archives of Neurology and Psychiatry, 37, 641–663.

    Google Scholar 

  • Darrow, C. W. (1937b). The equation of the galvanic skin reflex curve: I. The dynamics of reaction in relation to excitation-background. Journal of General Psychology, 16, 285–309.

    Google Scholar 

  • Darrow, C. W. (1964). The rationale for treating the change in galvanic skin response as a change in conductance. Psychophysiology, 1, 31–38.

    PubMed  Google Scholar 

  • Darrow, C. W., & Gullickson, G. R. (1970). The peripheral mechanism of the galvanic skin response. Psychophysiology, 6, 597–600.

    PubMed  Google Scholar 

  • Dawson, M. E., Schell, A. M., & Filion, D. L. (2007). The electrodermal system. In J. T. Cacioppo, L. G. Tassinary, & G. G. Berntson (Eds.), Handbook of psychophysiology (pp. 159–181). New York: Cambridge University Press.

    Google Scholar 

  • Ebbecke, U. (1951). Arbeitsweise der Schweißdrüsen und sudomotorische Reflexe bei unmittelbarer Beobachtung mit Lupenvergrößerung. Pflügers Archiv für die gesamte Physiologie, 253, 333–339.

    Google Scholar 

  • Edelberg, R. (1961). The relationship between the galvanic skin response, vasoconstriction, and tactile sensitivity. Journal of Experimental Psychology, 62, 187–195.

    PubMed  Google Scholar 

  • Edelberg, R. (1967). Electrical properties of the skin. In C. C. Brown (Ed.), Methods in psychophysiology (pp. 1–53). Baltimore: Williams & Wilkins.

    Google Scholar 

  • Edelberg, R. (1968). Biopotentials from the skin surface: The hydration effect. Annals of the New York Academy of Sciences, 148, 252–262.

    PubMed  Google Scholar 

  • Edelberg, R. (1971). Electrical properties of skin. In H. R. Elden (Ed.), A treatise of the skin (Biophysical properties of the skin, Vol. 1, pp. 519–551). New York: Wiley.

    Google Scholar 

  • Edelberg, R. (1972a). Electrical activity of the skin: Its measurement and uses in psychophysiology. In N. S. Greenfield & R. A. Sternbach (Eds.), Handbook of psychophysiology (pp. 367–418). New York: Holt, Rinehart, & Winston.

    Google Scholar 

  • Edelberg, R. (1973a). Mechanisms of electrodermal adaptations for locomotion, manipulation, or defense. In E. Stellar & J. M. Sprague (Eds.), Progress in physiological psychology (Vol. 5, pp. 155–209). New York: Academic.

    Google Scholar 

  • Edelberg, R. (1973b). The local electrical response of the skin to deformation. Journal of Applied Physiology, 34, 334–340.

    PubMed  Google Scholar 

  • Edelberg, R. (1983). The effects if initial levels of sweat duct filling and skin hydration on electrodermal response amplitude. Psychophysiology, 20, 550–557.

    PubMed  Google Scholar 

  • Edelberg, R. (1993). Electrodermal mechanisms: A critique of the two-effector hypothesis and a proposed replacement. In J.-C. Roy, W. Boucsein, D. C. Fowles, & J. H. Gruzelier (Eds.), Progress in electrodermal research (pp. 7–30). London: Plenum.

    Google Scholar 

  • Edelberg, R., & Wright, D. J. (1964). Two GSR effector organs and their stimulus specifity. Psychophysiology, 1, 39–47.

    PubMed  Google Scholar 

  • Ellis, R. A. (1968). Eccrine sweat glands: Electron microscopy; cytochemistry and anatomy. In O. Gans & G. K. Steigleder (Eds.), Handbuch der Haut – und Geschlechtskrankheiten (Normale und pathologische Anatomie der Haut, Vol. 1/1, pp. 224–266). Berlin: Springer.

    Google Scholar 

  • Féré, C. (1888). Note sur les modifications de la résistance électrique sous l’influence des excitations sensorielles et des émotions. Comptes Rendus des Séances de la Société de Biologie, 5, 217–219.

    Google Scholar 

  • Fisher, S. (1958). Body image and asymmetry of body reactivity. Journal of Abnormal and Social Psychology, 57, 292–298.

    Google Scholar 

  • Fitzgerald, M. J. T. (1961). Developmental changes in epidermal innervation. Journal of Anatomy, 95, 495–514.

    PubMed  Google Scholar 

  • Fowles, D. C. (1974). Mechanisms of electrodermal activity. In R. F. Thompson & M. M. Patterson (Eds.), Methods in physiological psychology (Bioelectric recording techniques, Part C: Receptor and effector processes, Vol. 1, pp. 231–271). New York: Academic.

    Google Scholar 

  • Fowles, D. C. (1986a). The eccrine system and electrodermal activity. In M. G. H. Coles, E. Donchin, & S. W. Porges (Eds.), Psychophysiology: Systems, processes, and applications (pp. 51–96). Amsterdam: Elsevier.

    Google Scholar 

  • Fowles, D. C., & Johnson, G. (1973). The influence of variations in electrolyte concentration on skin potential level and response amplitude. Biological Psychology, 1, 151–160.

    PubMed  Google Scholar 

  • Fowles, D. C., & Rosenberry, R. (1973). Effects of epidermal hydration on skin potential responses and levels. Psychophysiology, 10, 601–611.

    PubMed  Google Scholar 

  • Fowles, D. C., & Schneider, R. E. (1974). Effects of epidermal hydration on skin conductance responses and levels. Biological Psychology, 2, 67–77.

    PubMed  Google Scholar 

  • Fredrikson, M., Furmark, T., Olsson, M. T., Fischer, H., Andersson, J., & Langström, B. (1998). Functional neuoranatomical correlates of electrodermal activity: A positron emission tomographic study. Psychophysiology, 35, 179–185.

    PubMed  Google Scholar 

  • Fricke, M. (1932). Theory of electrolytic polarization. Philosophical Magazin and Journal of Science, 14, 310–318.

    Google Scholar 

  • Grice, K. A., & Verbov, J. (1977). Sweat glands and their disorders. In A. Rook (Ed.), Recent advances in dermatology (Vol. 4, pp. 155–198). New York: Churchill Livingstone.

    Google Scholar 

  • Grueninger, W. E., Kimble, D. P., Grueninger, J., & Levine, S. (1965). GSR and corticosteroid response in monkeys with frontal ablations. Neuropsychologia, 3, 205–216.

    Google Scholar 

  • Gruzelier, J. H. (1993). The laterality of electrodermal responses: A new perspective on individual differences in personality and psychopathology. In J.-C. Roy, W. Boucsein, C. D. Fowles, & J. H. Gruzelier (Eds.), Progress in Electrodermal Research (pp. 251–270). New York: Plenum.

    Google Scholar 

  • Hashimoto, K. (1978). The eccrine gland. In A. Jarrett (Ed.), The physiology and pathophysiology of the skin (The sweat glands, skin permeation, lymphatics, and the nails, Vol. 5, pp. 1543–1573). New York: Academic.

    Google Scholar 

  • Hermann, L., & Luchsinger, B. (1878). Über die Secretionsströme der Haut bei der Katze. Pflügers Archiv für die gesamte Physiologie, 19, 300–319.

    Google Scholar 

  • Herrmann, F., Ippen, H., Schaefer, H., & Stüttgen, G. (1973). Biochemie der Haut. Stuttgart: Thieme.

    Google Scholar 

  • Houdas, Y., & Ring, E. F. J. (1982). Human body temperature: Its measurement and regulation. New York: Plenum.

    Google Scholar 

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

    Google Scholar 

  • Isamat, F. (1961). Galvanic skin responses from stimulation of limbic cortex. Journal of Neurophysiology, 24, 176–181.

    PubMed  Google Scholar 

  • Jänig, W. (1979). Reciprocal reaction patterns of sympathetic subsystems with respect to various afferent inputs. In C. M. Brooks, K. Koizumi, & A. Sato (Eds.), Integrative functions of the autonomic nervous system. Amsterdam: Elsevier.

    Google Scholar 

  • Jänig, W. (1990). Functions of the sympathetic innervation of the skin. In A. D. Loewy & K. M. Spyer (Eds.), Central regulation of autonomic functions (pp. 334–347). New York: Oxford University Press.

    Google Scholar 

  • Jänig, W., Sundlöf, G., & Wallin, B. G. (1983). Discharge patterns of sympathetic neurons supplying skeletal muscle and skin in man and cat. Journal of the Autonomic Nervous System, 7, 239–256.

    PubMed  Google Scholar 

  • Jarrett, A. (1973a). The epidermis and its relations with the dermis. In A. Jarrett (Ed.), The physiology and pathophysiology of the skin (1, Vol. The epidermis). New York: Academic.

    Google Scholar 

  • Jarrett, A. (1973b). Normal epidermal keratinization. In A. Jarrett (Ed.), The physiology and pathophysiology of the skin (1, Vol. The epidermis). New York: Academic.

    Google Scholar 

  • Jarrett, A. (1980). Introduction: The permeability barrier. In A. Jarrett (Ed.), The physiology and pathophysiology of the skin (The mucous membranes, the action of vitamin A on the skin and mucous membranes, and transepidermal water loss, Vol. 6). New York: Academic.

    Google Scholar 

  • Jeje, A., & Koon, D. (1989). An analysis on the rates and regulation of insensible water loss through the eccrine sweat glands. Journal of Theoretical Biology, 141, 303–324.

    PubMed  Google Scholar 

  • Johnson, L. C., & Lubin, A. (1966). Spontaneous electrodermal activity during waking and sleeping. Psychophysiology, 3, 8–17.

    PubMed  Google Scholar 

  • Keller, P. (1963). Arbeitsphysiologie der Hornschicht in Grundzügen. In E. Schwarz, H. W. Spier, & G. Stüttgen (Eds.), Handbuch der Haut – und Geschlechtskrankheiten (Normale und pathologische Physiologie der Haut I, Vol. 1/3, pp. 36–89). Berlin: Springer.

    Google Scholar 

  • Kerassidis, S. (1994). Is palmar and plantar sweating thermoregulatory? Acta Physiologica Scandinavica, 152, 259–263.

    PubMed  Google Scholar 

  • Kimble, D. P., Bagshaw, M. H., & Pribram, K. H. (1965). The GSR of monkeys during orienting and habituation after selective partial ablations of the cingulate and frontal cortex. Neuropsychologia, 3, 121–128.

    Google Scholar 

  • Klaschka, F. (1979). Arbeitsphysiologie der Hornschicht in Grundzügen. In E. Schwarz, H. W. Spier, & G. Stüttgen (Eds.), Handbuch der Haut – und Geschlechtskrankheiten (1/4A, Vol. Normale und pathologische Physiologie der Haut II, pp. 153–261). Berlin: Springer.

    Google Scholar 

  • Knezevic, W., & Bajada, S. (1985). Peripheral autonomic surface potential: A quantitative technique for recording sympathetic conduction in man. Journal of the Neurological Sciences, 67, 239–251.

    PubMed  Google Scholar 

  • Kubota, Y., Sato, W., Murai, T., Toichi, M., Ikeda, A., & Sengoku, A. (2000). Emotional cognition without awareness after unilateral temporal lobectomy in humans. The Journal of Neuroscience, 97(RC 97), 1–5.

    Google Scholar 

  • Kuno, Y. (1956). Human perspiration. Springfield: Thomas.

    Google Scholar 

  • Kupfermann, J. (1985). Hypothalamus and limbic system II: Motivation. In E. R. Kandel & J. H. Schwartz (Eds.), Principles of neural science. Amsterdam: Elsevier.

    Google Scholar 

  • Lademann, J., Jacobi, U., Surber, C., Weigmann, H. J., & Fluhr, J. W. (2008). The tape stripping procedure – Evaluation of some critical parameters. European Journal of Pharmaceutics and Biopharmaceutics, 72, 317–323.

    PubMed  Google Scholar 

  • Ladpli, R., & Wang, G. H. (1960). Spontaneous variations of skin potentials in footpads of normal striatal and spinal cats. Journal of Neurophysiology, 23, 448–452.

    PubMed  Google Scholar 

  • Lang, H., Tuovinen, T., & Valleala, P. (1964). Amygdaloid afterdischarge and galvanic skin response. Electroencephalography and Clinical Neurophysiology, 16, 366–374.

    PubMed  Google Scholar 

  • Langworthy, O. R., & Richter, C. P. (1930). The influence of efferent cerebral pathways upon the sympathetic nervous system. Brain, 53, 178–193.

    Google Scholar 

  • LeDoux, J. (1996). The emotional brain: The mysterious underpinnings of emotional life. New York: Simon & Schuster.

    Google Scholar 

  • Lloyd, D. C. (1961). Action potential and secretory potential of sweat glands. Proceedings of the National Academy of Sciences USA, 47, 351–358.

    Google Scholar 

  • Love, T. J. (1980). Thermography as an indicator of blood perfusion. Annals of the New York Academy of Sciences, 335, 423–437.

    Google Scholar 

  • Löwenstein, W. R. (1956). Modulation of cutaneous receptors by sympathetic stimulation. The Journal of Physiology, 132, 40–60.

    Google Scholar 

  • Luria, A. R., & Homskaya, E. D. (1970). Frontal lobes and the regulation of arousal processes. In D. Mostofsky (Ed.), Attention: Contemporary theory and research (pp. 303–330). New York: Appleton Century Crofts.

    Google Scholar 

  • Lykken, D. T. (1968). Neuropsychology and psychophysiology in personality research. In E. F. Borgatta & W. W. Lambert (Eds.), Handbook of personality theory and research: Part 2. Psychophysiological techniques and personality research (pp. 413–509). Chicago: Rand McNally.

    Google Scholar 

  • Lykken, D. T. (1971). Square-wave analysis of skin impedance. Psychophysiology, 7, 262–275.

    Google Scholar 

  • Lykken, D. T., Miller, R. D., & Strahan, R. F. (1966). GSR and polarization capacity of skin. Psychonomic Science, 4, 355–356.

    Google Scholar 

  • Lykken, D. T., & Venables, P. H. (1971). Direct measurement of skin conductance: A proposal for standardization. Psychophysiology, 8, 656–672.

    PubMed  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, 1–8.

    PubMed  Google Scholar 

  • Marchionini, A., & Spier, H. W. (Eds.). (1963). Handbuch der Haut – und Geschlechtskrankheiten (Normale und pathologische Physiologie der Haut I, Vol. I/3). Berlin: Springer.

    Google Scholar 

  • Marsden, C. D. (1982). The mysterious motor function of the basal ganglia: The Robert Wartenberg Lecture. Neurology, 32, 514–539.

    PubMed  Google Scholar 

  • Martin, I., & Venables, P. H. (1966). Mechanisms of palmar skin resistance and skin potential. Psychological Bulletin, 65, 347–357.

    PubMed  Google Scholar 

  • Martin, I., & Venables, P. H. (Eds.). (1980). Techniques in Psychophysiology. Chicester: Wiley.

    Google Scholar 

  • Millington, P. F., & Wilkinson, R. (1983). Skin. Cambridge: University Press.

    Google Scholar 

  • Montagna, W., & Parakkal, P. F. (1974). The structure and function of skin. New York: Academic.

    Google Scholar 

  • Montagu, J. D., & Coles, E. M. (1966). Mechanism and measurement of the galvanic skin response. Psychological Bulletin, 65, 261–279.

    PubMed  Google Scholar 

  • Muthny, F. A. (1984). Elektrodermale Aktivität und palmare Schwitzaktivität als Biosignale der Haut in der psychophysiologischen Grundlagenforschung. Freiburg: Dreisam.

    Google Scholar 

  • Nagai, Y., Critchley, H. D., Featherstone, E., Trimble, M. R., & Dolan, R. J. (2004). Activity in ventromedial prefrontal cortex covaries with sympathetic skin conductance level: A physiological account of a “default mode” of brain function. NeuroImage, 22, 243–251.

    PubMed  Google Scholar 

  • Neumann, E., & Blanton, R. (1970). The early history of electrodermal research. Psychophysiology, 6, 453–475.

    PubMed  Google Scholar 

  • Nicolaidis, S., & Sivadjian, J. (1972). High-frequency pulsatile discharge of human sweat glands: Myoepithelial mechanism. Journal of Applied Physiology, 32, 86–90.

    PubMed  Google Scholar 

  • Niebauer, G. (1957). Der Aufbau des peripheren neurovegetativen Systems im Epidermal-Dermalbereich. Acta Neurovegetativa, 15, 109–123.

    PubMed  Google Scholar 

  • Nishiyama, T., Sugenoya, J., Matsumoto, T., Iwase, S., & Mano, T. (2001). Irregular activation of individual sweat glands in human sole observed by a videomicroscopy. Autonomic Neuroscience: Basic and Clinical, 88, 117–126.

    Google Scholar 

  • Obrist, P. A. (1963). Skin resistance levels and galvanic skin response: Unilateral differences. Science, 139, 227–228.

    PubMed  Google Scholar 

  • Odland, G. F. (1983). Structure of the skin. In L. A. Goldsmith (Ed.), Biochemistry and physiology of the skin (Vol. 1, pp. 3–63). New York: Oxford University Press.

    Google Scholar 

  • Orfanos, C. E. (1972). Feinstrukturelle Morphologie und Histopathologie der verhornenden Epidermis. Stuttgart: Thieme.

    Google Scholar 

  • Papez, J. W. (1937). A proposed mechanism of emotion. Archives of Neurology and Psychiatry, 38, 725–743.

    Google Scholar 

  • Pinkus, H. (1971). Embryology and anatomy of skin. In E. B. Helwig & F. K. Mostofi (Eds.), The skin (pp. 1–27). Baltimore: Williams & Wilkins.

    Google Scholar 

  • Prokasy, W. F., & Raskin, D. C. (Eds.). (1973). Electrodermal activity in psychological research. New York: Academic.

    Google Scholar 

  • Pugh, L. A., Oldroyd, C. A., Ray, T. S., & Clark, M. L. (1966). Muscular effort and electrodermal responses. Journal of Experimental Psychology, 71, 241–248.

    PubMed  Google Scholar 

  • 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 USA, 98, 676–682.

    Google Scholar 

  • Raine, A., & Lencz, T. (1993). Brain imaging research on electrodermal activity in humans. In J.-C. Roy, W. Boucsein, D. C. Fowles, & J. H. Gruzelier (Eds.), Progress in electrodermal research: From physiology to psychology (pp. 115–135). London: Plenum.

    Google Scholar 

  • Raine, A., Reynolds, G. P., & Sheard, C. (1991). Neuroanatomical correlates of skin conductance orienting in normal humans: A magnetic resonance imaging study. Psychophysiology, 28, 548–558.

    PubMed  Google Scholar 

  • Roberts, L. E., & Young, R. (1971). Electrodermal responses are independent of movement during aversive conditioning in rats, but heart rate is not. Journal of Comparative and Physiological Psychology, 77, 495–512.

    PubMed  Google Scholar 

  • Rothman, S. (1954). Physiology and biochemistry of the skin. Chicago: University of Chicago Press.

    Google Scholar 

  • Roy, J.-C., Boucsein, W., Fowles, D. C., & Gruzelier, J. H. (1993). Progress in Electrodermal Research: From Physiology to Psychology. New York: Plenum.

    Google Scholar 

  • Roy, J.-C., Delerm, B., & Granger, L. (1974). L’hibition bulbaire de l’activite electrodermale chez le chat. Electroencephalography and Clinical Neurophysiology, 37, 621–632.

    PubMed  Google Scholar 

  • Roy, J.-C., Sequeira, H., & Delerm, B. (1993). Neural control of electrodermal activity: Spinal and reticular mechanisms. In J.-C. Roy, W. Boucsein, C. D. Fowles, & J. H. Gruzelier (Eds.), Progress in Electrodermal Research (pp. 73–92). New York: Plenum.

    Google Scholar 

  • Roy, J.-C., Sequeira-Martinho, A. H., & Brochard, J. (1984). Pyramidal control of skin potential responses in the cat. Experimental Brain Research, 54, 283–288.

    Google Scholar 

  • Salter, D. C. (1979). Quantifying skin disease and healing in vivo using electrical impedance measurements. In P. Rolfe (Ed.), Non-invasive physiological measurements (Vol. 1, pp. 21–64). London: Academic.

    Google Scholar 

  • Salter, D. C. (1981). Alternating current electrical properties of human skin measured in vivo. In R. Marks & P. A. Payne (Eds.), Bioengineering and the skin (pp. 267–274). Lancaster: MTP.

    Google Scholar 

  • Sargent, F. I. (1962). Depression of sweating in man: So-called “Sweat Gland Fatigue”. In W. Mortagra, R. A. Ellis, & F. Silver (Eds.), Biology of the skin (Vol. III, pp. 163–212). Oxford: Pergamon.

    Google Scholar 

  • Sarkany, I., Shuster, S., & Stammers, M. C. (1965). Occlusion of the sweat pore by hydration. The British Journal of Dermatology, 77, 101–104.

    Google Scholar 

  • Sato, K. (1977). The physiology, pharmacology, and biochemistry of the eccrine sweat gland. Reviews of Physiology, Biochemistry and Pharmacology, 79, 51–131.

    PubMed  Google Scholar 

  • Sato, K. (1983). The physiology and pharmacology of the eccrine sweat gland. In L. A. Goldsmith (Ed.), Biochemistry and physiology of the skin (Vol. 1, pp. 596–641). New York: Oxford University Press.

    Google Scholar 

  • Sato, K., Kang, W. H., 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, 537–563.

    PubMed  Google Scholar 

  • Sato, K., & Sato, F. (1981). Pharmacologic responsiveness of isolated single eccrine sweat glands. The American Journal of Physiology, 240, R44–R51.

    PubMed  Google Scholar 

  • Schaefer, F., & Boucsein, W. (2000). Comparison of electrodermal constant voltage and constant current recording techniques using phase angle between alternating voltage and current. Psychophysiology, 37, 85–91.

    PubMed  Google Scholar 

  • Scheuplein, R. (1978). Site variations in diffusion and permeability. In A. Jarrett (Ed.), The physiology and pathophysiology of the skin (The sweat glands, skin permeation, lymphatics, and the nails, Vol. 5, pp. 1731–1752). New York: Academic.

    Google Scholar 

  • Schiffter, R., & Pohl, P. (1972). Zum Verlauf der absteigenden zentralen Sympathikusbahn. Archiv für Psychiatrie und Nervenkrankheiten, 216, 379–392.

    PubMed  Google Scholar 

  • Schiffter, R., & Schliack, H. (1968). Das sogenannte Geschmacksschwitzen. Fortschritte der Neurologie-Psychiatrie, 36, 262–274.

    Google Scholar 

  • Schliack, H., & Schiffter, R. (1979). Neurophysiologie und Pathophysiologie der Schweißsekretion. In E. Schwarz, H. W. Spier, & G. Stüttgen (Eds.), Handbuch der Haut – und Geschlechtskrankheite (1/4A, Vol. Normale und pathologische Physiologie der Haut II, pp. 349–458). Berlin: Springer.

    Google Scholar 

  • Schulz, I., Ullrich, K. J., Frömter, E., Holzgreve, H., Frick, A., & Hegel, U. (1965). Mikropunktion und elektrische Potentialmessung an Schweissdrüsen des Menschen. Pflügers Archiv für die gesamte Physiologie, 284, 360–372.

    Google Scholar 

  • Schwarz, E., Spier, H. W., & Stüttgen, G. (Eds.). (1979). Handbuch der Haut – und Geschlechtskrankheiten (Normale und pathologische Physiologie der Haut II, Vol. 1/4A). Berlin: Springer.

    Google Scholar 

  • Sequeira, H., & Roy, J.-C. (1993). Cortical and hypothalamo-limbic control of electrodermal responses. In J.-C. Roy, W. Boucsein, C. D. Fowles, & J. H. Gruzelier (Eds.), Progress in Electrodermal Research (pp. 93–114). New York: Plenum.

    Google Scholar 

  • Sequeira, H., & Roy, J. C. (1997). Neural control of electrodermal activity. In D. Jordan (Ed.), Central nervous control of autonomic function (pp. 259–293). Amsterdam: Harwood Academic.

    Google Scholar 

  • Sequeira-Martinho, H., Roy, J.-C., & Ba-M’hamed, S. (1986). Cortical and pyramidal stimulation elicit nonlateralized skin potential responses in the cat. Biological Psychology, 23, 81.

    Google Scholar 

  • Shaver, B. A., Brusilow, S. W., & Cooke, R. E. (1965). Electrophysiology of the sweat gland: Intraductal potential changes during secretion. Bulletin of the John Hopkins Hospital, 116, 100–109.

    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, 312–319.

    PubMed  Google Scholar 

  • Sinclair, D. (1973). Motor nerves and reflexes. In A. Jarrett (Ed.), The physiology and pathophysiology of the skin (The nerves and blood vessels, Vol. 2, pp. 475–508). New York: Academic.

    Google Scholar 

  • Spiegel, E. A., & Hunsicker, W. C. (1936). The conduction of cortical impulses to the autonomic system. The Journal of Nervous and Mental Disease, 83, 252–274.

    Google Scholar 

  • Stüttgen, G., & Forssmann, W. G. (1981). Pharmacology of the microvasculature of the skin. In G. Stüttgen, H. W. Spier, & E. Schwarz (Eds.), Handbuch der Haut – und Geschlechtskrankheiten (Normale und pathologische Physiologie der Haut III, Vol. 1/4B, pp. 379–540). Berlin: Springer.

    Google Scholar 

  • Tarchanoff, J. (1889). Decharges électriques dans la peau de l’homme sous l’influence de l’excitation des organes de sens et de differentes formes d’activité psychiqué. Comptes Rendus des Séances de la Société de Biologie, 41, 447–451.

    Google Scholar 

  • Tarchanoff, J. (1890). Über die galvanischen Erscheinungen an der Haut des Menschen bei Reizung der Sinnesorgane und bei verschiedenen Formen der psychischen Tätigkeit. Pflügers Archiv für die Gesamte Physiologie des Menschen und der Tiere, 46, 46–55.

    Google Scholar 

  • Tharp, M. D. (1983). Adrenergic receptors in the skin. In L. A. Goldsmith (Ed.), Biochemistry and physiology of the skin (Vol. 2, pp. 1210–1216). New York: Oxford University Press.

    Google Scholar 

  • Thews, G., Mutschler, E., & Vaupel, P. (1985). Human anatomy, physiology, and pathophysiology. Amsterdam: Elsevier.

    Google Scholar 

  • Thiele, F. A. J. (1981a). The functions of the atrichial (human) sweat gland. In G. Stüttgen, H. W. Spier, & E. Schwarz (Eds.), Handbuch der Haut – und Geschlechtskrankheiten (Normale und pathologische Physiologie der Haut III, Vol. 1/4B, pp. 2–121). Berlin: Springer.

    Google Scholar 

  • Thiele, F. A. J. (1981b). The sweat gland and the stratum corneum. In G. Stüttgen, H. W. Spier, & E. Schwarz (Eds.), Handbuch der Haut – und Geschlechtskrankheiten (Normale und pathologische Physiologie der Haut III, Vol. 1/4B, pp. 501–514). Berlin: Springer.

    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, 505–510.

    PubMed  Google Scholar 

  • Tranel, D. (2000). Electrodermal activity in cognitive neuroscience: Neuroanatomical and neuropsychological correlates. In R. D. Lane & L. Nadel (Eds.), Cognitive neuroscience of emotion (pp. 192–224). New York: Oxford University Press.

    Google Scholar 

  • Tranel, D., & Damasio, H. (1989). Intact electrodermal skin conductance responses after bilateral amygdala damage. Neuropsychologia, 27, 381–390.

    PubMed  Google Scholar 

  • Tranel, D., & Damasio, H. (1994). Neuroanatomical correlates of electrodermal skin conductance response. Psychophysiology, 31, 427–438.

    PubMed  Google Scholar 

  • Tregear, R. T. (1966). Physical functions of skin. London: Academic.

    Google Scholar 

  • Uno, H. (1977). Sympathetic innervation of the sweat glands and piloarrector muscles of macaques and human beings. The Journal of Investigative Dermatology, 69, 112–120.

    PubMed  Google Scholar 

  • Van de Staak, W. J. B. M. (1966). Experiences with the heated thermocouples method for measuring skin blood flow. Dermatologica, 132, 199–205.

    PubMed  Google Scholar 

  • Van der Valk, P. G. M., & Maibach, H. I. (1990). A functional study of the skin barrier to evaporative water loss by means of repeated cellophane-tape stripping. Clinical and Experimental Dermatology, 15, 180–182.

    PubMed  Google Scholar 

  • Venables, P. H., & Christie, M. J. (1973). Mechanisms, instrumentation, recording techniques, and quantification of responses. In W. F. Prokasy & D. C. Raskin (Eds.), Electrodermal activity in psychological research (pp. 1–124). New York: Academic.

    Google Scholar 

  • Venables, P. H., & Christie, M. J. (1980). Electrodermal activity. In I. Martin & P. H. Venables (Eds.), Techniques in psychophysiology (pp. 3–67). New York: Wiley.

    Google Scholar 

  • Veraguth, O. (1909). Das psychogalvanische Reflexphänomen. Berlin: Karger.

    Google Scholar 

  • Vigouroux, R. (1879). Sur le role de la résistance électrique de tissue dans l’électrodiagnostic. Comptes Rendus des Seances de la Société de Biologie, 31, 336–339.

    Google Scholar 

  • Wallin, B. G. (1981). Sympathetic nerve activity underlying electrodermal and cardiovascular reactions in man. Psychophysiology, 18, 470–476.

    PubMed  Google Scholar 

  • Wallin, B. G. (1992). Intraneural recordings of normal and abnormal sympathetic activity in man. In R. Bannister & C. J. Mathias (Eds.), Autonomic failure (pp. 359–377). Oxford: Medical.

    Google Scholar 

  • Wang, G. H. (1964). The neural control of sweating. Madison: University of Wisconsin Press.

    Google Scholar 

  • Wang, G. H., & Brown, V. W. (1956). Suprasegmental inhibition of an autonomic reflex. Journal of Neurophysiology, 19, 564–572.

    PubMed  Google Scholar 

  • Wang, G. H., & Lu, T. W. (1930). Galvanic skin reflex induced in the cat by stimulation of the motor area of the cerebral cortex. Chinese Journal of Physiology, 4, 303–326.

    Google Scholar 

  • Wechsler, D. (1925). The measurement of emotional reactions: Researches on the psychogalvanic reflexes. Archives of Psychology, 12, 1–181.

    Google Scholar 

  • Weiner, J. S., & Hellmann, K. (1960). The sweat glands. Biological Reviews, 35, 141–186.

    Google Scholar 

  • Weitkunat, R., Bührer, M., & Sparrer, B. (1990). Cortical initiation of phasic electrodermal activity. International Journal of Psychophysiology, 9, 303–314.

    PubMed  Google Scholar 

  • Wilcott, R. C. (1963). Effects of high environmental temperature on sweating and skin resistance. Journal of Comparative and Physiological Psychology, 56, 778–782.

    PubMed  Google Scholar 

  • Wilcott, R. C. (1965). A comparative study of the skin potential, skin resistance and sweating of the cat’s foot pad. Psychophysiology, 2, 62–71.

    PubMed  Google Scholar 

  • Wilcott, R. C. (1966). Adaptive value of arousal sweating and the epidermal mechanism related to skin potential and skin resistance. Psychophysiology, 2, 249–262.

    PubMed  Google Scholar 

  • Wilcott, R. C. (1969). Electrical stimulation of the anterior cortex and skin potential responses in the cat. Journal of Comparative and Physiological Psychology, 69, 465–472.

    PubMed  Google Scholar 

  • Wilcott, R. C., & Bradley, H. H. (1970). Low-frequency electrical stimulation of the cat’s anterior cortex and inhibition of skin potential changes. Journal of Comparative and Physiological Psychology, 72, 351–355.

    PubMed  Google Scholar 

  • Williams, L. M., Phillips, M. L., Brammer, M. J., Skerrett, D., Lagopoulos, J., Rennie, C., et al. (2001). Arousal dissociates amygdala and hippocampal fear responses: Evidence from simultaneous fMRI and skin conductance recording. NeuroImage, 14, 1070–1079.

    PubMed  Google Scholar 

  • Woodworth, R. S. & Schlosberg, H. (1954). Experimental psychology (3rd revised edn) New York: Holt, Rinehart, & Winston.

    Google Scholar 

  • Yamamoto, T., & Yamamoto, Y. (1976). Dielectric constant and resistivity of epidermal stratum corneum. Medical & Biological Engineering & Computing, 14, 494–500.

    Google Scholar 

  • Yamamoto, T., & Yamamoto, Y. (1981). Non-linear electrical properties of skin in the low frequency range. Medical & Biological Engineering & Computing, 19, 302–310.

    Google Scholar 

  • Yamamoto, Y., Yamamoto, T., Ohta, S., Uehara, T., Tahara, S., & Ishizuka, Y. (1978). The measurement principle for evaluating the performance of drugs and cosmetics by skin impedance. Medical & Biological Engineering & Computing, 16, 623–632.

    Google Scholar 

  • Yamazaki, K., Okamura, T., & Takasawa, N. (2001). Overt palmar surface sweating produces positive component of palmar skin potential responses. Japanese Journal of Physiological Psychology and Psychophysiology, 19, 1–6.

    Google Scholar 

  • Yokota, T., & Fujimori, B. (1964). Effects of brain-stem stimulation upon hippocampal electrical activity, somatomotor reflexes and autonomic functions. Electroencephalography and Clinical Neurophysiology, 16, 375–382.

    PubMed  Google Scholar 

  • Yokota, T., Sato, A., & Fujimori, B. (1963). Inhibition of sympathetic activity by stimulation of limbic systems. The Japanese Journal of Physiology, 13, 138–154.

    PubMed  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Wolfram Boucsein .

Rights and permissions

Reprints and permissions

Copyright information

© 2012 Springer Science+Business Media, LLC

About this chapter

Cite this chapter

Boucsein, W. (2012). Principles of Electrodermal Phenomena. In: Electrodermal Activity. Springer, Boston, MA. https://doi.org/10.1007/978-1-4614-1126-0_1

Download citation

Publish with us

Policies and ethics