Skip to main content

Methods of Electrodermal Recording

  • Chapter
  • First Online:
Electrodermal Activity

Abstract

The second chapter of the book discusses the different methods used for electrodermal recording. As mentioned in the introduction to Chap. 1, the observation of electrodermal phenomena requires only relatively basic equipment. As a consequence, a variety of recording methods have been proposed.

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.

    Although the respective EDL is built up of physical units with ratio scales, it is sometimes recommended that the EDL signal – like most psychophysiological variables – be treated as based only on interval scales (Levey, 1980; Stemmler, 1984), which would naturally also affect the performance of transformations (Sect. 2.3.3). However, EDR amp. obtained by the AC-coupled amplification cannot be treated as based on a ratio scale anyway.

  2. 2.

    The reading off and manual recording of the placement of the potentiometer R 3 in Fig. 2.3 cannot be recommended because of its error proneness.

  3. 3.

    Dirac impulses only exist in theory; in practice, a very narrow bandwidth square wave impulse with high amplitude is used. There may be danger of developing pain and erythema.

  4. 4.

    In a recently performed study for comparison of electrode creams for electrodermal recording, Tronstad, Johnsen, Grimnes, and Martinsen (2010; see also Sect. 2.2.2.5) applied constant voltage AC to contralateral palmar and abdominal sites. They observed an inverse EDR at abdominal sites which they interpreted as being due to an easier penetration of the hydrous components of the electrode gel into the sweat ducts, being facilitated by the thinner stratum corneum compared to palmar sites. They presumed that the mechanism of sweat reabsorption during recovery from sudomotor activity facilitates relatively moist and low-viscosity creams such as TD-246 (see Footnote 22 in Sect. 2.2.2.5) to penetrate into the ducts, thereby changing EDRs.

  5. 5.

    A comparison of different recording sites with SZL applied had been performed by Grimnes (1983; see Sect. 2.5.3.1).

  6. 6.

    Toyokura (1999) recorded SPRs following electric stimulation of the left median nerve at the wrist from 41 participants (9 females, 32 males; aged 22–60 years) from palmar and plantar sites in parallel, using reference electrodes on the nail of the index finger and the big toe. In general, SPRs at the soles had longer latencies and smaller amplitudes than those at the palms. These differences were larger than those between left- and right-recording sites. Furthermore, SPR waveforms (Sect. 2.3.1.2 “Amplitudes of Endosomatic Responses”) yielded not always congruent patterns on the palms and soles. However, these regional differences were not always reproducible.

  7. 7.

    No further specifications were provided for SC-recording in the paper. The electrodes were made from nickel plated brass and a stable 1.2 V reference was applied to the skin (Ouwerkerk, May 2011, personal communication). The device is also capable of storing data on onboard flash memory in off-line mode (de Vries, May 2011, personal communication). The data were sampled at 2 Hz, but actually, low-pass filtering was attained by taking eight samples in a row at 16 Hz which were then averaged (moving average). SCL and SCRs were evaluated, the latter by using the SCRGAUGE parameterization program from Kohlisch, published in the Appendix of Boucsein (1992). Because of the low quality of the SC signal, additional filtering was necessary.

  8. 8.

    At present, the use of such non-standard electrode sites cannot be recommended, since thorough comparisons of their electrodermal reactivity with the one of standard sites are still missing.

  9. 9.

    Silver plated 92%, nylon 8%, surface resistance <1Ω/sq and contact area of 3.5 cm2.

  10. 10.

    According to Venables and Christie (1980), no differences of skin potentials between abraded and non-abraded forearm sites were observable with children, so when children are used as participants, pretreatment will not be necessary.

  11. 11.

    For an improved handling, the overlaying rim of the adhesive collar can be attached horizontally to the rim of a table, to allow the experimenter using both hands for the electrode cream insertion.

  12. 12.

    The relative error due to seepage is dependent on electrode diameter. One millimeter of seepage increases the contact area to 2.25 times of its original size with 4 mm diameter miniature electrodes, but to only 1.13 times in case of 1 cm diameter electrodes are used (Venables & Christie, 1980).

  13. 13.

    Furthermore, to the present author’s experience, removing the second cover from the adhesive collar which is already fixed on the skin exerts some tension on the collar, which may lead to its detachment from the skin.

  14. 14.

    Histoacryl glue must be refrigerated. Experimenters must be also seriously cautioned, since careless use of histoacryl can result in eye damage. Also, acetone (which is also a cancer-causing suspect) might attack the rim of the electrode’s plastic chamber, thus roughen it and eventually damage its surface which is needed for the adherence to the collar.

  15. 15.

    Mastic adhesive is made from the Mastic pistachio tree. The German name is Mastix.

  16. 16.

    This corresponds to a 0.1 M solution of the monovalent NaCl (i.e., 0.58 g NaCl dissolved in 100 mL water).

  17. 17.

    The Beckman biopotential electrodes (see Footnote 18) show bias potentials of less than 250 μV and polarization potentials of less than 5 μV (Venables & Christie, 1980). Ag/AgCl electrodes can also be homemade, albeit unsintered (cf. Venables & Christie, 1973, p. 107), and good results are obtainable, but the manufacturing process is rather expensive, since pure silver (99.99%) would be necessary. Therefore, buying commercially available Ag/AgCl electrodes should be preferred. It should be noted that even with the use of so-called nonpolarizable electrodes there is still the possibility of counter e.m.f generation at the electrodes (Sect. 2.2.2.2).

  18. 18.

    The standard biopotential electrodes from Beckman Instruments have a contact area of 0.636 cm2. In Vivo-Metric-Systems electrodes, which are also sold alternatively together with a plug connection between the electrode and its wire, are of approximately the same size.

  19. 19.

    As to the present author’s experience, relatively long fluid storage may enable the saline penetrating into the plastic electrode chamber, which could result in corrosion of the contact between electrode and lead.

  20. 20.

    Sensor Medics recommended removing the deposit with dilute ammonium hydroxide. A five-to-one dilution with distilled water has been used with success (Vincent, July 1990, personal communication).

  21. 21.

    Recently, disposable EDA snap electrodes came into use, being distributed by BIOPAC Systems Inc. (http://www.biopac.com) and in Germany by med-NATIC (http://www.med-natic.de). These are Ag/AgCl electrodes of 2.5 × 4.5 cm size, filled with a 0.5% NaCl cream, with a foam backing and a stainless steel snap for electrode wires.

  22. 22.

    This is also true of the electrode cream “Synapse” made by Beckman Instruments which has been frequently used in EDA measurements. An analysis of this and other gels was made by Zipp, Hennemann, Grunwald, and Rohmert (1980), who found in “Synapse” a significant quantity of K and Cl ions in addition to Na ions. Grey and Smith (1984) reported that the Beckman cream has a NaCl concentration of 4.1 M, and contains glycerol, gum tragacanth, and 0.5% benzyl alcohol (as preservative). Beckman does no longer offer “Synapse,” and Unibase is also no longer available. The TD-246 electrode paste, which contains 0.5% saline in a neutral base, can be bought from Discount Disposables: http://www.discountdisposables.com/. According to the manufacturer, this cream “meets all the recommended specifications.” The cream is also distributed by Grass Technologies under the brand name EC33: http://www.grasstechnologies.com/ and as TD-246 in Germany by PAR Medizintechnik (http://www.PAR-Berlin.com).

  23. 23.

    Grey and Smith (1984) used a custom-made cream of 0.05 M NaCl solution in methylcellulose. However, they did not provide the correct formula of the ingredients they used (Clements, 1989).

  24. 24.

    Grey and Smith (1984) have published the following as being the ingredients of Unibase: cetyl and stearyl alcohols, soft paraffin, glycerol, and, as preservatives; 0.0015% propyl hydroxy-benzoate, sodium citrate, and sodium lauryl sulfate. The relative quantities are not provided. The water content is 63.4%. According to Grey and Smith, Unibase contains 0.028 mol/L Na ions. Back in the 1980s, the present author had this recipe chemically analyzed; the results were 0.07 mol/L Na and 0.045 mol/L Cl, the increased sodium being due to the Unibase itself (see Footnote 23). The cream is free from K and Ca ions (less than 0.01 g/kg) and has a nearly neutral pH value of 6.5. The analysis was performed by B. Neidhart, Institute for Industrial Physiology at the University of Dortmund, Germany.

  25. 25.

    Skin potential can be recorded with any EEG or EMG channel, which is a standard technique in neurology (Sect. 3.5.4).

  26. 26.

    Since Lykken and Venables (1971) recommended that skin conductance should be used as the appropriate unit for EDA measurement, constant current recording went out of use in most psychophysiological laboratories. However, because of the lower amplifier gain required, constant current methods are sometimes preferred in field applications (Sects. 2.1.1 and 2.6.2).

  27. 27.

    Though Venables and Christie (1980, p. 40) argued for the application of 0.5 V across each active site, following a suggestion by Edelberg (1967), their circuitry which underlies Fig. 2.10 provides a total of 0.5 V.

  28. 28.

    Note that this description is oriented towards paper recording (Sect. 2.2.4.1) which has come widely out of use. However, it is worth pondering such a procedure for a complete understanding of SC-recording, even if it is performed with A/D conversion of the recorded signal and subsequent computer evaluation (Sect. 2.2.4.2).

  29. 29.

    Simon and Homoth (1978) stated an error of 0.4% of the D/A converter’s output voltage. A circuit diagram of their voltage suppressor is given in their Fig. 1.2, together with a list of components in their Table 1.1.

  30. 30.

    The Vitaport ambulatory monitoring system is now distributed by TEMEC Instruments in Kerkrade, The Netherlands.

  31. 31.

    The Varioport ambulatory monitoring system is distributed by Becker Meditec in Karlsruhe, Germany.

  32. 32.

    Two pairs of standard Ag/AgCl electrodes were fixed on the same hand, one at standard thenar/hypothenar sites, the other one in 2–2.5 cm distance distal from the first ones (C and D in Fig. 2.6, Sect. 2.2.1.1).

  33. 33.

    Siemens 2-T VISION system for acquiring gradient-echo, echo-planar T*-weighted images with BOLD (blood oxygenation level dependent) contrast. Each volume comprised 48 × 3 mm axial scans with 3 mm inplane resolution, continuously acquired every 4.2 s. SC was recorded with Ag electrodes taped to the palmar surface of the left index and middle fingers, presumably without electrode cream.

  34. 34.

    Participants were scanned using a Siemens 1.5-T Magnetom VISION Plus system to acquire 64 T2*-weighted images depicting BOLD contrast for each stimulus of 3 s duration at 18 axial noncontinuous 6 mm thick planes (slices), parallel to the intracommissural line; sampling rate for the BOLD response (TR) = 3 s, TE = 40 ms, 128 × 128 matrix, interslice gap 0.6 mm.

  35. 35.

    The functional imaging data were acquired by a 1.5-T Siemens Symphony MRI-scanner with a Quantum gradient system. To measure the BOLD contrast, a T2*-weighted EPI (echo-planar imaging) sequence (TR = 2.5 s, TE = 60 ms, 64 × 64 matrix) was used. The volume contained 16 slices with a 5-mm slice thickness (no gap). The slices were acquired interleaved in ascending order. Artifacts stemming from recording (Sect. 2.2.5.1) were reduced with automatic 1D de-noising using wavelets. The threshold selection rule was a heuristic variant of Stein’s Unbiased Risk (Matlab R12).

  36. 36.

    Functional and structural MRI scans were obtained using a GE Signa 1.5-T scanner. The participant’s neck and head were stabilized within foam padding within a brain-specific RF head coil. An EPI pulse sequence was used for collecting functional data (TR = 2.5 s, TE = 30 ms, FOV = 24 cm, 64 × 64 matrix). The entire brain was covered in 20 or 21 slices in the sagittal plane, resulting in voxel dimensions of 3.75 × 3.75 × 7 mm. Additionally, a high-resolution fast 3D T1-weighted structural image (TE = 6 ms, FOV = 24 cm, 256 × 256 × 124 voxels of 1.9 × 1.9 × 2 mm) was obtained as anatomical reference. SC was recorded by AgCl electrodes from the index and middle fingers of the participant’s left hand by means of a commercially available recording system which was placed outside the magnet room. Analog SC signals were recorded at 10 Hz and passed to an A/D converter. No more SC-recording details were provided.

  37. 37.

    The conversion rate depends on the A/D converter used in the computer system. Thus, Venables and Christie (1980) recommend 20 Hz, while Foerster (1984) used 16 Hz, which makes no difference in accuracy with respect to the phasic changes that occur in the electrodermal signal.

  38. 38.

    After gentle removal of their electrodes being for 24 h in place, Tronstad et al. (2010) took photographs of the skin areas that were blindly assessed by a dermatologist for skin irritation by different electrode creams. Only some signs of very faint erythema were observed, the degree of which was always lower than on the skin areas under the adhesive tape.

  39. 39.

    Recently, omitting of an electrode cream became rather common in EDA recording devices for applications outside the laboratory (Sects. 2.2.3.4 and 2.2.6.3). However, it must be kept in mind that such a system will be unstable for an unknown period of time, since the humidity built up by sweat under dry electrodes will cause a drift towards an increase of skin conductance.

  40. 40.

    For a multiple-electrode recording technique using electrical impedance spectroscopy, see Footnote 50, Sect. 1.4.3.3.

  41. 41.

    Small deviations from this standard methodology will be mentioned in the text, larger ones in footnotes.

  42. 42.

    Nishiyama, Sugenoya, Matsumoto, Iwase, and Mano (2001), in their study described in Footnote 17, Sect. 1.3.2.1, observed that sudomotor bursts as recorded by microneurography were followed by SPRs with latencies of 1.33 ± 0.33 s.

  43. 43.

    The convention for graphical representation is, as in the neurophysiological tradition of EEG recording, “negative up” (Venables & Christie, 1980).

  44. 44.

    An illustration of two subsequent EDRs, from both of which all EDR parameters can be obtained without extrapolation because the recovery of the first EDR goes beyond half of its amplitude is shown in Fig. 18.3 of Boucsein (2005).

  45. 45.

    A declaration of the signal-to-noise ratio, which is obvious in audio devices, is often lacking in descriptions of polygraph amplifiers.

  46. 46.

    The first derivate has been used by Biro and Stukovsy (1993) in an evaluation of paper-recorded SRRs to 1 kHz, 100 dB, 550 ms tones from 300 male participants. Besides the SRR amp. evaluated from the original curve, several amplitude, time, and reaction shape parameters were exploited from the first derivate. A factor analysis of all parameters revealed a response shape factor, an amplitude factor and a latency factor, which altogether accounted for 84.6% of the total variance.

  47. 47.

    Thom (1988) used a criterion of 10% instead, because the application of the 1% criterion is difficult if numerous electrodermal fluctuations appear.

  48. 48.

    A is measured in mV for SP, in μS for SC and SY, and in kΩ for SR and SZ.

  49. 49.

    The curve was kindly made available by F. Foerster, University of Freiburg, Germany.

  50. 50.

    Refined mathematical modeling of EDR curves had been performed by Hunt (1977), who developed an equation based on overlapping Gaussian distributions to fit the course of SRRs, and by Schneider (1987). Schneider fitted a three-compartment model to the recorded SC curve (personal communication) which includes the physical properties of the duct filling, the active membrane response in the duct walls, and the corneal hydration (Sect. 1.4.2). Schneider could show that a typical SCR can be modeled by assuming a roughly triangular input signal and choosing as an impulse response a sum of two exponentials with time constants of approximately 2 and 20 s, respectively. More recently, several authors came forward with similar proposals for the mathematical modeling of overlapping SCRs which are generated by short ISIs (Sect. 2.3.1.5).

  51. 51.

    The half-life period is defined by these authors – in contrast to the nomenclature used in section 2.3.1.3 under “Recovery Parameters” – as the time the curve remains over the half amplitude (T/2) in Fig. 2.20. It therefore includes the time characteristics of the ascent as well.

  52. 52.

    In their mathematical model, sudomotor nerve discharges were regarded as Dirac impulses (see Sect. 1.4.1.4). According to the convolution theorem, the Fourier transformed SC – SCL signal time series equals the product of the Fourier transformed nerve discharges with the Fourier transformed response function (see equation (3) in Bach et al., 2010). This means that the sudomotor nerve discharge frequency will have its greatest impact on the SC spectral power if it matches the peak frequencies of the response function. If the spectral power of the response function is known, it will be possible to recover the sudomotor nerve firing frequency according to equation (5) in Bach et al. However, unknown noise and response variability may prevent such an approach.

  53. 53.

    The appropriateness of this assumption is questionable. In general, mathematical models have not yet overcome the general problem that EDRs reveal so different shapes. Even if the model fits the majority of the empirically observed EDR recoveries, there is still a remainder of forms which differ so much from the identified standard that models may fail in evaluating them adequately.

  54. 54.

    As already demonstrated by Boucsein (1988, 1992), a sufficient approximation of a typical SCR can be attained by a combination of two e-functions for the descent and another e-function for the ascent (see Fig. 2.18, Sect. 2.3.1.3).

  55. 55.

    This is an assumption which is not in line with the variability in SCR shape observed by various researchers (cf. Benedek & Kaernbach, 2010) (see also Footnote 53).

  56. 56.

    In the present author’s view, this assumption does not meet what can be observed in the majority of EDRs, which do not return to the SCL prior to their onset, even in case of nonsuperimposed SCRs (Sect. 2.3.1.3 “Recovery Parameters”). Bach et al. (2009) artificially attained such a return to zero by applying high-pass filtering, thus removing slow components of the SCR which may reflect important processes resulting from moistening the corneum (Sect. 1.4.2.3).

  57. 57.

    A nonstandard recording method has been applied, using a 10 V source in series with a 13.2 MΩ resistor over dry Ag/AgCl electrodes of 10 mm diameter. SC data were sampled at 32 Hz and 24 bit A/D conversion. As an amplitude criterion, 0.01 μS was applied. To improve distributional characteristics, data were logarithmically transformed (Sect. 2.3.3.3).

  58. 58.

    An increase of SCR amp. with increasing ISI length was also found by Breska, Maoz, and Ben-Shakhar (2011), who compared ISIs ranging from 16 to 24 s (mean 20 s) with ISIs shortened by 50% in a within-subjects design with 36 participants (19 females, 17 males).

  59. 59.

    The same is true for the EDL values which are calculated from the decoupled AC curve of overlapping impulses (Sects. 2.1.3 and 2.2.4.2).

  60. 60.

    Edelberg (1967) proposed using the frequency of changes in the EDL within a certain time span instead of the NS.EDR freq. as an indicator of arousal, a method which has not yet been applied to the present author’s knowledge.

  61. 61.

    Following the same rationale, impedance and admittance values might be related to the electrode area. Such kind of transformation is not common in endosomatic recording.

  62. 62.

    This has been demonstrated by Boucsein, Baltissen, and Euler (1984) using both recording methods in parallel during the application of 2-s white noise stimuli with intensities between 60 and 110 dB.

  63. 63.

    ALS scores can be standardized not only intraindividually over the different EDRs, but also interindividually for each response over all participants. In the latter case, both score sequences X and Y are calculated using the EDRs of all participants to the same stimulus.

  64. 64.

    The computer program described in the Appendix of this book allows for a display of the respiratory signal together with the EDA curve to be evaluated.

  65. 65.

    Several authors used the term “EDR magnitude” instead of “EDR amplitude” (Sect. 2.3.1.2). Therefore, care should be taken in ascertaining just which method of evaluation was used in the respective publications.

  66. 66.

    Mathematically this is a type of missing data treatment, as nonappearing EDRs, or EDRs which remain below an amplitude criterion (Sect. 2.3.1.2 “Choice of Amplitude Criteria”) are taken into account in evaluation. The EDR magnitude (more precisely, the mean EDA magnitude) is calculated by dividing the sum of the evaluated EDR amp. by the number of occasions in which EDRs might have been expected.

  67. 67.

    This is understandable if one notes that Venables and Christie give 21°C as the correct temperature for a European laboratory, which is in the present author’s experience somewhat too low. Instead, in his own laboratory, the present author maintained a constant temperature of 23°C and 50% relative humidity.

  68. 68.

    The authors reported that they had to raise the laboratory temperature to an unusually high 30°C, since at lower temperatures the Mauritians displayed hardly any EDRs.

  69. 69.

    In addition, functional and morphological changes in eccrine sweat glands have been observed in vitro and in vivo during heat acclimatization in three patas monkeys by Sato, Owen, Matthes, Sato, and Gisolfi (1990).

  70. 70.

    Recorded as SR by means of a modified Wheatstone bridge (Fig. 2.3, Sect. 2.1.3), converted into log resistance values (Sect. 2.3.3.3). The recording device allowed for up to 20 active electrodes being connected in rapid succession. Recording sites were: fingertips, middle and proximal phalanges of the fingers, several palmar (including thenar and hypothenar) sites, the dorsal side of the hand, the wrist, several volar, and one dorsal point on the forearm.

  71. 71.

    This study was performed with electrodes that were mechanically pressed to the skin.

  72. 72.

    The dry electrodes used in this study may have contributed to this inconsistency.

  73. 73.

    One interesting result should be mentioned here: Christie and Venables (1971) observed negative correlations between the BSPL (Sect. 2.3.2.1) and the T-wave amplitude (TWA) in the ECG, from investigating 21 male participants lying down (r = −0.70) and from another 15 participants in a sitting position (r = −0.61). The authors suggest the extracellular potassium ionic concentration as being the cause for both an increased TWA and an increased negativity of the BSPL. Furedy and Heslegrave (1983) suggested that the TWA constitutes an index of excitatory sympathetic activity. Since the EDA is a valid index of sympathetic excitation, a high correlation between TWA and EDA could have been expected.

  74. 74.

    In this preparation, the active sweat glands appear as holes, since the plastic film does not attach to water (i.e., sweat on top of the open pore), which can be counted. This method had been discussed as a cheap alternative to recording tonic EDA, labeled “palmar sweat index” (Turpin & Clements, 1993) but was rather seldom used to date.

  75. 75.

    EDA was measured as SRL with a 2.54 cm2 dry silver disk electrode and transformed into SCLs.

  76. 76.

    EDA was measured as SR through Ag/AgCl sponge electrodes of 1 cm diameter with an “inert” electrolyte using 40 μA current and a Wheatstone bridge, being transformed to SC values.

  77. 77.

    In their comparison of eight postmenopausal older females (52–62 years) with eight younger females (20–30 years) during exercise under dry heat, Anderson and Kenney (1987) observed a lower sweating rate in the older group, which reflected a diminished output per heat-activated sweat gland rather than a decrease in the number of sweat glands recruited.

  78. 78.

    Recorded with pure silver spiral electrodes that were chlorided electrolytically.

  79. 79.

    Recorded as SR with 46 μA from palm vs. forearm with 1 cm2 Ag/AgCl electrodes held in place by an elastic band, using cellulose sponge holders soaked with saline as electrolyte.

  80. 80.

    To the present author’s experience, electrodermal nonreactivity in elderly study participants may be a result of too low ambient temperatures and hence generally reduced sweat gland activity (Sect. 2.4.1.1). Therefore, care should be taken for creating comfortable climatic conditions for elderly participants, e.g., by raising the laboratory temperature several degrees or covering their body with a light blanket.

  81. 81.

    Recorded with standard methodology with the use of a Wheatstone bridge from the volar surfaces of the distal phalanges of the first and second fingers of the nonpreferred hand. The recording interval was 1–5 s after stimulus onset (1 kHz, 75 dB tones at irregular intervals between 30 and 90 s); the amplitude criterion was 0.05 μS.

  82. 82.

    Using 2 cm diameter zinc electrodes, embedded in a plastic cup, which was filled with 1% zinc sulfate in agar paste. A 40 μA current was applied to measure SR, the SRR amp was determined and transformed to log SC change.

  83. 83.

    Recorded with silver electrodes covered with an AgCl layer attached by means of flexible wires to palmar, plantar, and calf regions, using a constant voltage of 1.35 V.

  84. 84.

    In their study on seasonal effects on the SRL described in Sect. 2.2.1.1, Neumann (1968) revealed a less differentiated pattern in children between 6 and 11 years compared to adults.

  85. 85.

    With 1.0 V CC, using a Wheatstone bridge, 7.5 mm diameter Ag/AgCl electrodes thenar/hypothenar from the nondominant hand.

  86. 86.

    Unipolar recording taken from the palm and the sole with Ag/AgCl electrodes of 7 mm diameter, filled with Beckman electrode cream enriched with additional salt. The inactive electrode was attached to the dorsum of the respective hand and foot, and another one on the forearm as control site.

  87. 87.

    Recorded by zinc/zinc sulfate saline electrodes from palmar sites, transformed to log values.

  88. 88.

    Method not reported in detail, presumably unipolar.

  89. 89.

    With zinc electrodes and zinc sulfate electrode cream from palmar sites.

  90. 90.

    Recorded with standard methodology, however using Beckman cream, as SR and transformed to SC. Differences appeared in both SCL and NS.SCR freq.

  91. 91.

    Maximum SCL reached during the presentation of five slides, recorded with K–Y gel, the type of electrodes not being mentioned.

  92. 92.

    Gender differences in electrodermal conditioning have also be observed in dependence of the person’s gender who expressed emotions on slides used as CSs. Mazurski, Bond, Siddle, and Lovibond (1996) presented to 52 females and 35 males angry faces as CS+ and neutral faces as CS− in a picture-shock conditioning paradigm, crossing the expressing person’s gender with the gender of participants. Male participants showed larger SCR amp. (recorded with standard methodology) to expressions of males than did female participants, whereas the respective SCR amp. did not differ in female participants. Therefore, the “preparedness” hypothesis of conditioning (Sect. 3.1.2.1 “UCR Diminution, Preception and Preparedness”) could only be confirmed for male participants.

  93. 93.

    Increasing thoracic pressure through trying an exhalation with mouth and nose shut, following deep breathing.

  94. 94.

    Recorded with 7 mm diameter Ag/AgCl electrodes, 0.068 M NaCl cream and two 0.2 V constant voltage amplifiers on both hands simultaneously.

  95. 95.

    Gender differences have also been observed in EDA recordings during the presentation of emotion-inducing films (Kring & Gordon, 1998; see Footnote 158 in Sect. 3.2.2.1) and in pictures with emotional content (Bradley, Codispoti, Sabatinelli, & Lang, 2001). In the first study of the latter authors performed with 50 females and 45 males, a marginal gender main effect was obtained for the SC peak, indicating that larger SC changes were observed when males viewed opposite-sex erotic pictures, compared with females.

  96. 96.

    Recorded with CV from 10 mm diameter Ag/AgCl electrodes, filled with 17% NaCl gel; no further specifications provided.

  97. 97.

    Recordings were performed with 0.2 V “constant current” from 7 mm diameter Ag/AgCl electrodes filled with 0.068 M NaCl gel, attached to the thenar/hypothenar eminences of the right hand.

  98. 98.

    Described only as taken from the fingers and expressed in μS.

  99. 99.

    For more information on sweat gland activity in postmenopausal females, see Footnote 77 in Sect. 2.4.3.1.

  100. 100.

    Using a 0.5 V constant voltage and 1.5 cm diameter Ag/AgCl ambulatory monitoring electrodes, filed with 0.05 M KCl Unibase-glycol paste. Ambulatory recording was performed with a Medilog tape recorder.

  101. 101.

    It is sometimes said that Chinese people do not have sweat glands. This erroneous statement may result from sweat gland activity being normally lower in Asiatic people, due to their sweat glands being smaller as compared to those of Caucasians.

  102. 102.

    Recorded with stainless steel disc electrodes of 9.5 mm diameter, filled with so-called Cambridge cream, by means of a tissue resistance monitor providing 8 Hz square wave and a constant current of 20 μA.

  103. 103.

    For further discussion of experimenter’s ethnic group on participant’s physiological reactions, see Venables and Christie (1973).

  104. 104.

    Measured palmar/dorsal at the dominant hand by Ag/AgCl electrodes and Beckman electrode cream, using “a constant direct current of 20 μV” (which should be presumably μA).

  105. 105.

    Barabasz (1970) observed significant more EDA changes during an imagination task in 19 African-Americans compared to 20 Caucasians.

  106. 106.

    Method of EDA measurement as used by the Johnson group (see Footnote 131, Sect. 3.2.1.3).

  107. 107.

    Obtained with a so-called Fels Dermohmeter and zinc electrodes from the palms, and in one group from the plantar arch.

  108. 108.

    Recorded with standard methodology, using constant current of 10 μA/cm2.

  109. 109.

    Jorgenson et al. (1988) in their study described in the Sect. 2.4.3.1, revealed by using the finger-sweat print method that the number of active hypothenar sweat glands per 1/4 cm² was considerably higher in African-American compared to White newborns, being reversed in children of mean age 7–8 years and in adults of about 25 years, with some gender differences as well (cf., Jorgenson et al., Table 3).

  110. 110.

    Measured with standard methodology (using Beckman cream) as skin resistance (but with 20 μA constant current), transformed to SC.

  111. 111.

    These results are not really convincing, since the Caucasian Bedouin sample participants were partly gathered by the police and were being moved, both actions may raise SCL markedly. In another study, Kugelmass and Lieblich (1968) observed higher SCL and lower electrodermal reactivity in Bedouin samples as compared to Israeli samples.

  112. 112.

    Presentation of 16 combinations of a tone (1 kHz, 68 dB, 35 s) with a 110 dB white noise of 0.5 s duration, the length of which could be shortened by pressing a button.

  113. 113.

    Measured by Ag/AgCl sponge electrodes from the right palm.

  114. 114.

    Recorded with CV, 8 mm diameter Ag/AgCl electrodes with an isotonic NaCl paste from the participant’s left index and middle fingers, scoring the maximum SCR during 1–4 s after stimulus onset.

  115. 115.

    A review of reliabilities of different EDA parameters including various investigations was provided by Freixa i Baqué (1982).

  116. 116.

    Since phasic SP measures are so much dependent upon experimental as well as recording conditions, reporting typical distribution parameters and reliabilities will not be attempted.

  117. 117.

    They used tin electrodes with a contact surface area of 0.72 cm2, probably without electrolytes.

  118. 118.

    Recorded with disc electrodes of 10 mm in diameter from the volar and dorsal side of the hand and the foot. SPRs were evoked by various stimuli at irregular intervals: deep inspiration, single auditory clicks and electrical stimulation of the median nerve at the wrist, the tibial nerve at the ankle, and the supraorbital nerve at the forehead.

  119. 119.

    The method of determining the BSPL as the minimal obtainable SPL has already been discussed in Sect. 2.3.2.1.

  120. 120.

    They used liquid electrolytes (KCl/agar) and calomel (mercury chloride) electrodes.

  121. 121.

    They used the earlobe as an inactive site for SP recordings, and dry silver electrodes with 3.8 cm2 surfaces for SR recordings.

  122. 122.

    Venables’ team used KCl-based electrode cream. Although sweat contains by far more NaCl than KCl, the difference between the use of those monovalent ions is given little importance in the literature on methodology (Sect. 2.2.2.5).

  123. 123.

    Since the SCL scores are also dependent upon the type and concentration of the electrolytes, the distribution data from the study of Venables’ team cannot easily be generalized, as they used a KCl cream that is not commonly used for SC measurements.

  124. 124.

    Based on the same data set, Fahrenberg, Schneider, and Safian (1987) reported short-term stabilities between r = 0.03 and 0.27 under resting conditions and between r = 0.05 and 0.32 under performance conditions for the NS.SCR freq. (amplitude criterion = 0.3 μS). A short-term stability of r = 0.57 was reported within a 30-min session from initial to final resting.

  125. 125.

    For mathematical solutions for overlapping EDRs, see Sect. 2.3.1.5.

  126. 126.

    Three weeks being between the first and second recordings and between the third and fourth recordings, and 6 weeks being between the second and third recordings.

  127. 127.

    The authors used standard methodology, with electrodes of 12 mm in diameter, but transformed the SR scores into SC units before calculating nonspecific responses.

  128. 128.

    In this study, an unusually high current of 70 μA was used.

  129. 129.

    Recorded by means of zinc electrodes with a contact surface area of 0.32 cm2, zinc sulfate as electrolyte and a constant current method with 3.0 μA; results being transformed into conductance units, and the SCR amp. being square root transformed (Sect. 2.3.3.3).

  130. 130.

    Recording was performed with constant current using concentric electrodes with an internal diameter of 5 mm and an external diameter of 0.6–1 cm, using 0.05 M KCl cream on an agar base, from the index and middle fingers of the left hand.

  131. 131.

    Recovery measured in percent of amplitude decrease 2 s after the point of maximum deflection.

  132. 132.

    They used stainless steel electrodes with 2 cm diameter, at a distance of 2 cm on the volar middle of the underarm, fastened with rubber bands to filter paper soaked with a NaCl solution.

  133. 133.

    When a baseline score (a) and a response score (b) are uncorrelated, the correlation of the reactivity measure (a − b) and the baseline score (a) cannot equal zero because they have a common term (Myrtek & Foerster, 1986). For many physiological variables, (a) and (b) are not totally independent from each other, which leads to differently high correlations between (a) and (b), and therefore to a differently large a(a − b) effect.

  134. 134.

    Palmar recording with zinc sulfate cream, at 40 μA.

  135. 135.

    As a consequence, one should rather use the term “concept of initial values” instead of the term “law.”

  136. 136.

    A thorough discussion of such corrections which are based either on the use of transformations or on regression techniques is found in Levey (1980, p. 619 ff.).

  137. 137.

    See (1.6e) in Sect. 1.4.1.2. To avoid confusion, the parameter x is used instead of R in (1.22) in Sect. 1.4.3.1.

  138. 138.

    When the resistance R 2 (or the conductance G 2) of the epidermis are also regarded as variable, Equations (2.25c), respectively (2.26c) become more complicated as they must be differentiated for a second variable. However, in this case, different level dependencies will also result.

  139. 139.

    In spite of Lykken and Venables gave a clear recommendation for constant voltage methods as early as in 1971 and the SPR Publication recommendations adhered to this 10 years later (Fowles et al., 1981), both types of methods are still in use as can be inferred from various studies cited in the present book.

  140. 140.

    This discussion has been taken up by Catania et al. (1980) to explain the dependency of finding age-related differences in electrodermal reactivity upon the method of measurement (Sect. 2.4.3.1).

  141. 141.

    The points raised by Edelberg (1967, p. 25 f.) on technical control of the current density with constant voltage systems are not further discussed here since they are taken care of in modern equipment (Sect. 2.2.4).

  142. 142.

    Zinc electrodes of 21 mm diameter together with zinc sulfate as electrode cream were used.

  143. 143.

    Constant current recordings might have an advantage over constant voltage ones during long-term recordings, as Edelberg (1967) revealed (Sect. 2.2.6.1).

  144. 144.

    With 0.6 cm2 Beckman Ag/AgCl electrodes, Hellige isotonic electrode cream (Sect. 2.2.2.5), 0.5 V constant voltage and 10 μA/cm2 constant current. The participants received 30 acoustic stimuli at intensities varying between 60 and 110 dB.

  145. 145.

    In a study where 20 participants were presented 10 tones of 50 dB each, and additionally white quadrangles as stimuli. With the constant current measurements, Barry (1981) used polarizable electrodes and non-isotonic cream, while the constant voltage measurements were performed using standard methodology.

  146. 146.

    In the present author’s view, other attempts which take advantage of computerized EDR evaluation have not as yet demonstrated their superiority to traditional procedures within a theoretically underpinned context, such as discarding the theoretically founded FIR/SIR evaluation of EDRs within classical conditioning (Sect. 3.1.2.1 “Recent Developments in EDR Conditioning”). At least, all attempts to modify standard EDA evaluation procedures should include a diligent empirical comparison with traditional evaluation methods.

References

  • Ahlberg, J. H., Nilson, E. N., & Walsh, J. L. (1967). The theories of splines and their applications. New York: Academic.

    Google Scholar 

  • Alexander, D. M., Trengove, C., Johnston, P., Cooper, T., August, J. P., & Gordon, E. (2005). Separating individual skin conductance responses in a short interstimulus-interval paradigm. Journal of Neuroscience Methods, 146, 116–123.

    Article  PubMed  Google Scholar 

  • Almasi, J. J., & Schmitt, O. H. (1974). Automated measurement of bioelectrical impedance at very low frequencies. Computers and Biomedical Research, 7, 449–456.

    Article  PubMed  Google Scholar 

  • Anderson, R. K., & Kenney, W. L. (1987). Effect of age on heat-activated sweat gland density and flow during exercise in dry heat. Journal of Applied Physiology, 63, 1089–1094.

    PubMed  Google Scholar 

  • Andresen, B. (1987). Differentielle Psychophysiologie valenzkonträrer Aktivierungsdimensionen. Frankfurt: Peter Lang.

    Google Scholar 

  • Andrew, W., & Winston-Salem, N. C. (1966). Structural alternations with aging in the nervous system. Journal of Chronic Disease, 3, 575–596.

    Article  Google Scholar 

  • Arena, J. G., Blanchard, E. B., Andrasik, F., Cotch, P. A., & Myers, P. E. (1983). Reliability of psychophysiological assessment. Behavior Research and Therapy, 21, 447–460.

    Article  Google Scholar 

  • Asso, D., & Braier, J. R. (1982). Changes with the menstrual cycle in psychophysiological and self-report measures of activation. Biological Psychology, 15, 95–107.

    Article  PubMed  Google Scholar 

  • Bach, D. R., Flandin, G., Friston, K. J., & Dolan, R. J. (2009). Time-series analysis for rapid event-related skin conductance responses. Journal of Neuroscience Methods, 184, 224–234.

    Article  PubMed  Google Scholar 

  • Bach, D. R., Friston, K. J., & Dolan, R. J. (2010). Analytic measures for quantification of arousal from spontaneous skin conductance fluctuations. International Journal of Psychophysiology, 76, 52–55.

    Article  PubMed  Google Scholar 

  • Balin, A. K., & Pratt, L. A. (1989). Physiological consequences of human skin aging. Cutis, 43, 431–436.

    PubMed  Google Scholar 

  • Baltissen, R. (1983). Psychische und somatische Reaktionen auf affektive visuelle Reize bei jungen und alten Personen. Unpublished Doctoral Dissertation, Düsseldorf.

    Google Scholar 

  • Freixa i Baqué, E. (1982). Reliability of electrodermal measures: A compilation. Biological Psychology, 14, 219–229.

    Google Scholar 

  • Barabasz, A. F. (1970). Galvanic skin response and test anxiety among Negros and Caucasians. Child Study Journal, 1, 33–35.

    Google Scholar 

  • Barry, R. J. (1981). Comparability of EDA effects obtained with constant-current skin resistance and constant-voltage skin conductance methods. Physiological Psychology, 9, 325–328.

    Google Scholar 

  • Becker-Carus, C., & Schwarz, E. (1981). Differentielle Unterschiede psychophysiologischer Aktivierungsverläufe und Kurzzeitgedächtnisleistungen in Abhängigkeit von Persönlichkeitskriterien. In W. Janke (Ed.), Beiträge zur Methodik in der differentiellen, diagnostischen und klinischen Psychologie (pp. 87–103). Königstein: Anton Hain.

    Google Scholar 

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

    PubMed  Google Scholar 

  • Benjamin, L. S. (1967). Facts and artifacts in using analysis of covariance to “undo” the law of initial values. Psychophysiology, 4, 187–206.

    Article  PubMed  Google Scholar 

  • Ben-Shakhar, G. (1985). Standardization within individuals: A simple method to neutralize individual differences in skin conductance. Psychophysiology, 22, 292–299.

    Article  PubMed  Google Scholar 

  • Ben-Shakhar, G., Lieblich, I., & Kugelmass, S. (1975). Detection of information and GSR habituation: An attempt to derive detection efficiency from two habituation curves. Psychophysiology, 12, 283–288.

    Article  Google Scholar 

  • Berardesca, E., de Rigal, J., Leveque, J. L., & Maibach, H. I. (1991). In vivo biophysical characterization of skin physiological differences in races. Dermatologica, 182, 89–93.

    Article  PubMed  Google Scholar 

  • Bernstein, A. S. (1965). Race and examiner as significant influences on basal skin impedance. Journal of Personality and Social Psychology, 1, 346–349.

    Article  Google Scholar 

  • Besthorn, D., Schellberg, D., Pfleger, W., & Gasser, T. (1989). Using variance as a tonic SCR parameter. Journal of Psychophysiology, 3, 419–424.

    Google Scholar 

  • Biro, V., & Stukovsky, R. (1993). Components of the GSR curve in adult males. Studia Psychologica, 35, 111–117.

    Google Scholar 

  • Bitterman, M. E., & Holtzman, W. H. (1952). Conditioning and extinction of the galvanic skin response as a function of anxiety. Journal of Abnormal and Social Psychology, 47, 615–623.

    Article  Google Scholar 

  • Blank, I. H., & Finesinger, J. E. (1946). Electrical resistance of the skin. Archives of Neurology and Psychiatry, 56, 544–557.

    PubMed  Google Scholar 

  • Blecker, C. R., Kirsch, P., Schaefer, F., & Vaitl, D. (2001). Skin conductance measurement during fMRT scans: A methodology study. Psychophysiology, 38, S27.

    Google Scholar 

  • Block, J. D., & Bridger, W. H. (1962). The law of initial value in psychophysiology: A reformulation in terms of experimental and theoretical considerations. Annals of the New York Academy of Sciences, 98, 1229–1241.

    Article  PubMed  Google Scholar 

  • Botwinick, J., & Kornetsky, C. (1960). Age differences in the acquisition and extinction of the GSR. Journal of Gerontology, 15, 83–84.

    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. (2001). Physiologische Grundlagen und Meßmethoden der dermalen Aktivität. In F. Rösler (Ed.), Enzyklopädie der Psychologie, Bereich Psychophysiologie, Band 1: Grundlagen und Methoden der Psychophysiologie (pp. 551–623). Göttingen: Hogrefe.

    Google Scholar 

  • Boucsein, W. (2005). Electrodermal measurement. In N. Stanton, A. Hedge, K. Brookhuis, E. Salas, & H. Hendrick (Eds.), Handbook of human factors and ergonomic methods (pp. 18–1–18–8). CRC: Boca Raton.

    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.

    Article  PubMed  Google Scholar 

  • Boucsein, W., & Hoffmann, G. (1979). A direct comparison of the skin conductance and skin resistance methods. Psychophysiology, 16, 66–70.

    Article  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.

    Article  PubMed  Google Scholar 

  • Boucsein, W., Schaefer, F., & Sommer, T. (2001). Electrodermal long-term monitoring in everyday life. In J. Fahrenberg & M. Myrtek (Eds.), Progress in ambulatory assessment (pp. 549–560). Göttingen: Hogrefe.

    Google Scholar 

  • Boucsein, W., & Thum, M. (1996). Multivariate psychophysiological analysis of stress-strain processes under different break schedules during computer work. In J. Fahrenberg & M. Myrtek (Eds.), Ambulatory assessment (pp. 305–313). Seattle: Hogrefe & Huber.

    Google Scholar 

  • Boucsein, W., & Thum, M. (1997). Design of work/rest schedules for computer work based on psychophysiological recovery measures. International Journal of Industrial Ergonomics, 20, 51–57.

    Article  Google Scholar 

  • Boucsein, W., et al. (2012). Publication recommendations for electrodermal measurements: An update. Psychophysiology, 49 (in preparation).

    Google Scholar 

  • Bradley, M. M., Codispoti, M., Sabatinelli, D., & Lang, P. J. (2001). Emotion and motivation II: Sex differences in picture processing. Emotion, 1, 300–319.

    Article  PubMed  Google Scholar 

  • Breska, A., Maoz, K., & Ben-Shakhar, G. (2011). Interstimulus intervals for skin conductance response measurement. Psychophysiology, 48, 437–440.

    Article  PubMed  Google Scholar 

  • Brown, C. C. (1972). Instruments in psychophysiology. In N. S. Greenfield & R. A. Sternbach (Eds.), Handbook of psychophysiology (pp. 159–195). New York: Holt, Rinehart, & Winston.

    Google Scholar 

  • Bull, R. H. C., & Gale, A. (1971). The relationships between some measures of the galvanic skin response. Psychonomic Science, 25, 293–294.

    Google Scholar 

  • Bull, R. H. C., & Gale, A. (1973). The reliability of and interrelationships between various measures of electrodermal activity. Journal of Experimental Research in Personality, 6, 300–306.

    Google Scholar 

  • Bull, R., & Gale, A. (1974). Does the law of initial value apply to the galvanic skin response? Biological Psychology, 1, 213–227.

    Article  PubMed  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.

    Article  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.

    Article  Google Scholar 

  • Burstein, K. R., Fenz, W. D., Bergeron, J., & Epstein, S. (1965). A comparison of skin potential and skin resistance responses as measures of emotional responsivity. Psychophysiology, 2, 14–24.

    Article  PubMed  Google Scholar 

  • Burton, C. E., David, R. M., Portnoy, W. M., & Akers, L. A. (1974). The application of bode analysis to skin impedance. Psychophysiology, 11, 517–525.

    Article  PubMed  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.

    Article  PubMed  Google Scholar 

  • Catania, J. J., Thompson, L. W., Michalewski, H. A., & Bowman, T. E. (1980). Comparisons of sweat gland counts, electrodermal activity, and habituation behavior in young and old groups of subjects. Psychophysiology, 17, 146–152.

    Article  PubMed  Google Scholar 

  • Christie, M. J., & Venables, P. H. (1971). Basal palmar skin potential and the electrocardiogram T-wave. Psychophysiology, 8, 779–786.

    Article  PubMed  Google Scholar 

  • Christie, M. J., & Venables, P. H. (1972). Site, state, and subject characteristics of palmar skin potential levels. Psychophysiology, 9, 645–649.

    Article  PubMed  Google Scholar 

  • Clements, K. (1989). The use of purpose-made electrode gels in the measurement of electrodermal activity: A correction to Grey and Smith (1984). Psychophysiology, 26, 495.

    Article  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.

    Article  PubMed  Google Scholar 

  • Corah, N. L., & Stern, J. A. (1963). Stability and adaptation of some measures of electrodermal activity in children. Journal of Experimental Psychology, 65, 80–85.

    Article  PubMed  Google Scholar 

  • Cort, J., Hayworth, J., Little, B., Lobstein, T., McBrearty, E., Reszetniak, S., et al. (1978). The relationship between the amplitude and the recovery half-time of the skin conductance response. Biological Psychology, 6, 309–311.

    Article  Google Scholar 

  • Crider, A., & Lunn, R. (1971). Electrodermal lability as a personality dimension. Journal of Experimental Research in Personality, 5, 145–150.

    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 

  • Curzi-Dascalova, L., Pajot, N., & Dreyfus-Brisac, C. (1973). Spontaneous skin potential responses in sleeping infants between 24 and 41 weeks of conceptional age. Psychophysiology, 10, 478–487.

    Article  PubMed  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.

    Article  Google Scholar 

  • Davis, T., Love, B. C., & Maddox, W. T. (2009). Anticipatory emotions in decision tasks: Covert markers of value or attentional processes? Cognition, 112, 195–200.

    Article  PubMed  Google Scholar 

  • De Jongh, G. J. (1981). Porosity of human skin in vivo assessed via water loss, carbon dioxide loss and electrical impedance for healthy volunteers, atopic and psoriatic patients. Current Problems in Dermatology, 9, 83–101.

    PubMed  Google Scholar 

  • Deltombe, T., Hanson, P., Jamart, J., & Clérin, M. (1998). The influence of skin temperature on latency and amplitude of the sympathetic skin response in normal subjects. Muscle & Nerve, 21, 34–39.

    Article  Google Scholar 

  • Docter, R., & Friedman, L. F. (1966). Thirty-day stability of spontaneous galvanic skin responses in man. Psychophysiology, 2, 311–315.

    Article  Google Scholar 

  • Edelberg, R. (1964). Independence of galvanic skin response amplitude and sweat production. The Journal of Investigative Dermatology, 42, 443–448.

    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. (1970). The information content of the recovery limb of the electrodermal response. Psychophysiology, 6, 527–539.

    Article  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. (1972b). Electrodermal recovery rate, goal-orientation, and aversion. Psychophysiology, 9, 512–520.

    Article  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.

    Chapter  Google Scholar 

  • Edelberg, R., Greiner, T., & Burch, N. R. (1960). Some membrane properties of the effector in the galvanic skin response. Journal of Applied Physiology, 15, 691–696.

    PubMed  Google Scholar 

  • Edelberg, R., & Muller, M. (1981). Prior activity as a determinant of electrodermal recovery rate. Psychophysiology, 18, 17–25.

    Article  PubMed  Google Scholar 

  • Eisdorfer, C. (1978). Psychophysiological and cognitive studies in the aged. In G. Usdin & D. J. Hofling (Eds.), Aging: The process and the people (pp. 96–128). New York: Brunner/Mazel.

    Google Scholar 

  • Eisdorfer, C., Doerr, H. O., & Follette, W. (1980). Electrodermal reactivity: An analysis by age and sex. Journal of Human Stress, 6, 39–42.

    Article  PubMed  Google Scholar 

  • Eisenstein, E. M., Bonheim, P., & Eisenstein, D. (1995). Habituation of the galvanic skin response to tone as a function of age. Brain Research Bulletin, 37, 343–350.

    Article  PubMed  Google Scholar 

  • El-Sheikh, M. (2007). Children’s skin conductance level and reactivity: Are these measures stable over time and across tasks? Developmental Psychobiology, 49, 180–186.

    Article  PubMed  Google Scholar 

  • Faber, S. (1977). Methodische Probleme bei Hautwiderstandsmessungen. Biomedizinische Technik, 22, 393–394.

    Article  Google Scholar 

  • Faber, S. (1980). Hautleitfähigkeitsuntersuchungen als Methode in der Arbeitswissenschaft. Düsseldorf: Fortschritt-Berichte der VDI-Zeitschriften.

    Google Scholar 

  • Fahrenberg, J., & Foerster, F. (1982). Covariation and consistency of activation parameters. Biological Psychology, 15, 151–169.

    Article  PubMed  Google Scholar 

  • Fahrenberg, J., Foerster, F., Schneider, H. J., Müller, W., & Myrtek, M. (1984). Aktivierungsforschung im Labor-Feld-Vergleich. München: Minerva.

    Google Scholar 

  • Fahrenberg, J., & Myrtek, M. (1967). Zur Methodik der Verlaufsanalyse: Ausgangswerte, Reaktionsgrößen (Reaktivität) und Verlaufswerte. Psychologische Beiträge, 10, 58–77.

    Google Scholar 

  • Fahrenberg, J., & Myrtek, M. (Eds.). (1996). Ambulatory assessment. Seattle: Hogrefe & Huber.

    Google Scholar 

  • Fahrenberg, J., & Myrtek, M. (Eds.). (2001). Progress in Ambulatory Assessment. Göttingen: Hogrefe.

    Google Scholar 

  • Fahrenberg, J., Schneider, H.-J., & Safian, P. (1987). Psychophysiological assessments in a repeated-measurement design extending over a one-year interval: Trends and stability. Biological Psychology, 24, 49–66.

    Article  PubMed  Google Scholar 

  • Fahrenberg, J., Walschburger, P., Foerster, F., Myrtek, M., & Müller, W. (1979). Psychophysiologische Aktivierungsforschung: Ein Beitrag zu den Grundlagen der multivariaten Emotions – und Stress-Theorie. München: Minerva.

    Google Scholar 

  • Fenske, N. A., & Lober, C. W. (1986). Structural and functional changes of normal aging skin. Journal of the American Academy of Dermatology, 15, 571–585.

    Article  PubMed  Google Scholar 

  • Fisher, L. E., & Kotses, H. (1973). Race differences and experimenter race effect in galvanic skin response. Psychophysiology, 10, 578–582.

    Article  PubMed  Google Scholar 

  • Fisher, L. E., & Winkel, M. H. (1979). Time of quarter effect: An uncontrolled variable in electrodermal research. Psychophysiology, 16, 158–163.

    Article  PubMed  Google Scholar 

  • Fletcher, R. P., Venables, P. H., & Mitchell, D. A. (1982). Estimation of half from quarter recovery time of SCR. Psychophysiology, 19, 115–116.

    Article  PubMed  Google Scholar 

  • Foerster, F. (1984). Computerprogramme zur Biosignalanalyse. Berlin: Springer.

    Google Scholar 

  • Forbes, T. W. (1964). Problems in measurement of electrodermal phenomena – choice of method and phenomena – potential, impedance, resistance. Psychophysiology, 1, 26–30.

    Article  PubMed  Google Scholar 

  • Forbes, T. W., & Landis, C. (1935). The limiting A. C. frequency for the exhibition of the galvanic skin (“psychogalvanic”) response. The Journal of General Psychology, 13, 188–193.

    Article  Google Scholar 

  • Foulds, I. S., & Barker, A. T. (1983). Human skin battery potentials and their possible role in wound healing. British Journal of Dermatology, 109, 515–522.

    Article  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., Christie, M. J., Edelberg, R., Grings, W. W., Lykken, D. T., & Venables, P. H. (1981). Publication recommendations for electrodermal measurements. Psychophysiology, 18, 232–239.

    Article  PubMed  Google Scholar 

  • Fowles, D. C., Kochanska, G., & Murray, K. (2000). Electrodermal activity and temperament in preschool children. Psychophysiology, 37, 777–787.

    Article  PubMed  Google Scholar 

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

    Article  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.

    Article  PubMed  Google Scholar 

  • Fowles, D. C., & Schneider, R. E. (1978). Electrolyte medium effects on measurements of palmar skin potential. Psychophysiology, 15, 474–482.

    Article  PubMed  Google Scholar 

  • Francini, F., Zoppi, M., Maresca, M., & Procacci, P. (1979). Skin potential and EMG changes induced by cutaneous electrical stimulation. Applied Neurophysiology, 42, 113–124.

    PubMed  Google Scholar 

  • Fredrikson, M. (1986). Racial differences in cardiovascular reactivity to mental stress in essential hypertension. Journal of Hypertension, 4, 325–331.

    Article  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.

    Article  PubMed  Google Scholar 

  • Freedman, R. R., Woodward, S., & Norton, D. A. M. (1992). Laboratory and ambulatory monitoring of menopausal hot flushes: Comparison of symptomatic and asymptomatic women. Journal of Psychophysiology, 6, 162–166.

    Google Scholar 

  • Fried, R. (1982). On-line analysis of the GSR. The Pavlovian Journal of Biological Science, 17, 89–94.

    PubMed  Google Scholar 

  • Furchtgott, E., & Busemeyer, J. K. (1979). Heart rate and skin conductance during cognitive processes as a function of age. Journal of Gerontology, 34, 183–190.

    PubMed  Google Scholar 

  • Furedy, J. J., & Heslegrave, R. J. (1983). A consideration of recent criticisms of the t-wave amplitude index of myocardial sympathetic activity. Psychophysiology, 20, 204–211.

    Article  PubMed  Google Scholar 

  • Galbrecht, C. R., Dykman, R. A., Reese, W. G., & Suzuki, T. (1965). Intrasession adaptation and intersession extinction of the components of the orienting response. Journal of Experimental Psychology, 70, 585–597.

    Article  PubMed  Google Scholar 

  • Gao, Y., Raine, A., Dawson, M. E., Venables, P. H., & Mednick, S. A. (2007). Development of skin conductance orienting, habituation, and reorienting from ages 3 to 8 years: A longitudinal latent growth curve analysis. Psychophysiology, 44, 855–863.

    Article  PubMed  Google Scholar 

  • Garwood, M., Engel, B. T., & Kusterer, J. P. (1981). Skin potential level: Age and epidermal hydration effects. Journal of Gerontology, 36, 7–13.

    PubMed  Google Scholar 

  • Garwood, M., Engel, B. T., & Quilter, R. E. (1979). Age differences in the effect of epidermal hydration on electrodermal activity. Psychophysiology, 16, 311–317.

    Article  PubMed  Google Scholar 

  • Gavazzeni, J., Wiens, S., & Fischer, H. (2008). Age effects to negative arousal differ for self-report and electrodermal activity. Psychophysiology, 45, 148–151.

    PubMed  Google Scholar 

  • Gaviria, B., Coyne, L., & Thetford, P. E. (1969). Correlation of skin potential and skin resistance measures. Psychophysiology, 5, 465–477.

    Article  PubMed  Google Scholar 

  • Germana, J. (1968). Rate of habituation and the law of initial values. Psychophysiology, 5, 31–36.

    Article  PubMed  Google Scholar 

  • Goldstein, J. M., Jerram, M., Abbs, B., Whitfield-Gabrieli, S., & Makris, N. (2010). Sex differences in stress response circuitry activation dependent on female hormonal cycle. Journal of Neuroscience, 30, 431–438.

    Article  PubMed  Google Scholar 

  • Goldstein, J. M., Jerram, M., Poldrack, R., Ahern, T., Kennedy, D. N., Seidman, L. J., et al. (2005). Hormonal cycle modulates arousal circuitry in women using functional magnetic resonance imaging. Journal of Neuroscience, 25, 9309–9316.

    Article  PubMed  Google Scholar 

  • Gómez-Amor, J., Martínez-Selva, J. M., Román, F., Zamora, S., & Sastre, J. F. (1990). Electrodermal activity, hormonal levels and subjective experience during the menstrual cycle. Biological Psychology, 30, 125–139.

    Article  PubMed  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, 551–557.

    Article  PubMed  Google Scholar 

  • Grimnes, S. (1982). Psychogalvanic reflex and changes in electrical parameters of dry skin. Medical & Biological Engineering & Computing, 20, 734–740.

    Article  Google Scholar 

  • Grimnes, S. (1983). Impedance measurement of individual skin surface electrodes. Medical & Biological Engineering & Computing, 21, 750–755.

    Article  Google Scholar 

  • Grimnes, S., Jabbari, A., Martinsen, O. G., & Tronstad, C. (2010). Electrodermal activity by DC potential and AC conductance measured simultaneously at the same skin site. Skin Research and Technology, 17, 26–34.

    Article  Google Scholar 

  • Grings, W. W. (1974). Recording of electrodermal phenomena. In R. F. Thompson & M. M. Patterson (Eds.), Methods in physiological psychology (Bioelectric recording techniques, Part C: Receptor and effector processes, Vol. 1, pp. 273–296). New York: Academic.

    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, 77–82.

    Article  PubMed  Google Scholar 

  • Gruzelier, J. H., & Venables, P. H. (1972). Skin conductance orienting activity in a heterogeneous sample of schizophrenics: Possible evidence of limbic dysfunction. The Journal of Nervous and Mental Disease, 155, 277–287.

    Article  PubMed  Google Scholar 

  • Hagfors, C. (1964). Beiträge zur Meßtheorie der hautgalvanischen Reaktion. Psychologische Beiträge, 7, 517–538.

    Google Scholar 

  • Hare, R. D., Wood, K., Britain, S., & Frazelle, J. (1971). Autonomic responses to affective visual stimulation: Sex differences. Journal of Experimental Research in Personality, 5, 14–22.

    Google Scholar 

  • Hettema, J. M., Annas, P., Neale, M. C., Kendler, K. S., & Fredrikson, M. (2003). A twin study of the genetics of fear conditioning. Archives of General Psychiatry, 60, 702–708.

    Article  PubMed  Google Scholar 

  • Hinton, J., O’Neill, M., Dishman, J., & Webster, S. (1979). Electrodermal indices of public offending and recidivism. Biological Psychology, 9, 297–309.

    Article  PubMed  Google Scholar 

  • Hölzl, R., Wilhelm, H., Lutzenberger, W., & Schandry, R. (1975). Galvanic skin response: Some methodological considerations on measurement, habituation, and classical conditioning. Archiv für Psychologie (Archives of Psychology), 127, 1–22.

    Google Scholar 

  • Hord, D. J., Johnson, L. C., & Lubin, A. (1964). Differential effect of the law of initial value (LIV) on autonomic variables. Psychophysiology, 1, 79–87.

    Article  PubMed  Google Scholar 

  • Hoyt, C. J. (1941). Note on a simplified method of computing test reliability. Educational and Psychological Measurement, 1, 93–95.

    Article  Google Scholar 

  • Hunt, D. P. (1977). A mathematical model of a simple human galvanic skin response upon its rate topography. Bulletin of the Psychonomic Society, 10, 149–151.

    Google Scholar 

  • Hupka, R. B., & Levinger, G. (1967). Within subject correspondence between skin conductance and skin potential under conditions of activity and passivity. Psychophysiology, 4, 161–167.

    Article  PubMed  Google Scholar 

  • Hustmyer, F. E., & Burdick, J. A. (1965). Consistency and test-retest reliability of spontaneous autonomic nervous system activity and eye movements. Perceptual and Motor Skills, 20, 1225–1228.

    Article  PubMed  Google Scholar 

  • Hygge, S., & Hugdahl, K. (1985). Skin conductance recordings and the NaCl concentration of the electrolyte. Psychophysiology, 22, 365–367.

    Article  PubMed  Google Scholar 

  • Iacono, W. G., Lykken, D. T., Haroian, K. P., Peloquin, L. J., Valentine, R. H., & Tuason, V. B. (1984). Electrodermal activity in euthymic patients with affective disorders: One-year retest stability and the effects of stimulus intensity and significance. Journal of Abnormal Psychology, 93, 304–311.

    Article  PubMed  Google Scholar 

  • Janes, C. L. (1982). Electrodermal recovery and stimulus significance. Psychophysiology, 19, 129–135.

    Article  PubMed  Google Scholar 

  • Janes, C. L., Hesselbrock, V., & Stern, J. A. (1978). Parental psychopathology, age, and race as related to electrodermal activity of children. Psychophysiology, 15, 24–34.

    Article  PubMed  Google Scholar 

  • Janes, C. L., Strock, B. D., Weeks, D. G., & Worland, J. (1985). The effect of stimulus significance on skin conductance recovery. Psychophysiology, 22, 138–146.

    Article  PubMed  Google Scholar 

  • Janes, C. L., Worland, J., & Stern, J. (1976). Skin potential and vasomotor responsiveness of black and white children. Psychophysiology, 13, 523–527.

    Article  PubMed  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 

  • Johnson, L. C. (1963). Some attributes of spontaneous autonomic activity. Journal of Comparative and Physiological Psychology, 56, 415–422.

    Article  Google Scholar 

  • Johnson, L. C., & Corah, N. L. (1963). Racial differences in the skin. Science, 139, 766–767.

    Article  PubMed  Google Scholar 

  • Johnson, L. C., & Landon, M. M. (1965). Eccrine sweat gland activity and racial differences in resting skin conductance. Psychophysiology, 1, 322–329.

    Article  PubMed  Google Scholar 

  • Johnson, L. C., & Lubin, A. (1972). On planning psychophysiological experiments: Design, measurement, and analysis. In N. S. Greenfield & R. A. Sternbach (Eds.), Handbook of psychophysiology (pp. 125–158). New York: Holt, Rinehart, & Winston.

    Google Scholar 

  • Jones, B. E., & Ayres, J. J. B. (1966). Significance and reliability of shock-induced changes in basal skin conductance. Psychophysiology, 2, 322–326.

    Article  Google Scholar 

  • Jorgenson, R. J., Salinas, C. F., Dowben, J. S., & St. John, D. L. (1988). A population study on the density of palmar sweat pores. Birth Defects, 24, 51–63.

    PubMed  Google Scholar 

  • Juniper, K., & Dykman, R. (1967). Skin resistance, sweat gland counts, salivary flow, and gastric secretion: Age, race, and sex differences, and intercorrelations. Psychophysiology, 4, 216–222.

    Article  PubMed  Google Scholar 

  • Kaelbling, R., King, F. A., Achenbach, K., Branson, R., & Pasamanick, B. (1960). Reliability of autonomic responses. Psychological Reports, 6, 143–163.

    Article  Google Scholar 

  • Katkin, E. S. (1965). Relationship between manifest anxiety and two indices of autonomic response to stress. Journal of Personality and Social Psychology, 2, 324–333.

    Article  Google Scholar 

  • Kaye, H. (1964). Skin conductance in the human neonate. Child Development, 35, 1297–1305.

    PubMed  Google Scholar 

  • Ketterer, M. W., & Smith, B. D. (1977). Bilateral electrodermal activity, lateralized cerebral processing and sex. Psychophysiology, 14, 513–516.

    Article  PubMed  Google Scholar 

  • Kimmel, H. D., & Hill, F. A. (1961). A comparison of two electrodermal measures of response to stress. Journal of Comparative and Physiological Psychology, 54, 395–397.

    Article  PubMed  Google Scholar 

  • Kimmel, H. D., & Kimmel, E. (1965). Sex differences in adaptation of the GSR under repeated applications of a visual stimulus. Journal of Experimental Psychology, 70, 536–537.

    Article  PubMed  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.

    Article  PubMed  Google Scholar 

  • Koglbauer, I., Kallus K. W., Braunstingl, R., & Boucsein, W. (2011). Recovery training in simulator improves performance and psychophysiological state of pilots during simulated and real VFR flight. The International Journal of Aviation Psychology, 21 (in press).

    Google Scholar 

  • Köhler, T., Vögele, C., & Weber, D. (1989). Die Zahl der aktiven Schweißdrüsen (PSI, Palmar Sweat Index) als psychologischer Parameter. Zeitschrift für Experimentelle und Angewandte Psychologie, 24, 89–100.

    Google Scholar 

  • Kopacz, G. M., & Smith, B. D. (1971). Sex differences in skin conductance measures as a function of shock threat. Psychophysiology, 8, 293–303.

    Article  PubMed  Google Scholar 

  • Korol, B., Bergfeld, G., & McLaughlin, L. J. (1975). Skin color and autonomic nervous system measures. Psychology and Behavior, 14, 575–578.

    Google Scholar 

  • Korol, B., & Kane, R. (1978). An examination of the relationship between race, skin color and a series of autonomic nervous system measures. The Pavlovian Journal of Biological Science, 13, 121–132.

    PubMed  Google Scholar 

  • Kring, A. M., & Gordon, A. H. (1998). Sex differences in emotion: Expression, experience, and physiology. Journal of Personality and Social Psychology, 74, 686–703.

    Article  PubMed  Google Scholar 

  • Kryspin, J. (1965). The phoreographical determination of the electrical properties of human skin. The Journal of Investigative Dermatology, 44, 227–229.

    PubMed  Google Scholar 

  • Kugelmass, S., & Lieblich, I. (1968). Relation between ethnic origin and GSR reactivity in psychophysiological detection. Journal of Applied Psychology, 52, 158–162.

    Article  PubMed  Google Scholar 

  • Lacey, J. I. (1956). The evaluation of autonomic responses: Toward a general solution. Annals of the New York Academy of Sciences, 67, 125–163.

    Article  PubMed  Google Scholar 

  • Lader, M. H. (1970). The unit of quantification of the G.S.R. Journal of Psychosomatic Research, 14, 109–110.

    Article  PubMed  Google Scholar 

  • Lathrop, R. G. (1964). Measurement of analog sequential dependencies. Human Factors, 6, 233–239.

    PubMed  Google Scholar 

  • Lawler, J. C., Davis, M. J., & Griffith, E. C. (1960). Electrical characteristics of the skin: The impedance of the surface sheath and deep tissues. The Journal of Investigative Dermatology, 34, 301–308.

    PubMed  Google Scholar 

  • Levander, S. E., Schalling, D. S., Lidberg, L., Bartfai, A., & Lidberg, Y. (1980). Skin conductance recovery time and personality in a group of criminals. Psychophysiology, 17, 105–111.

    Article  PubMed  Google Scholar 

  • Leveque, J. L., Corcuff, P., de Rigal, J., & Agache, P. (1984). In vivo studies of the evolution of physical properties of the human skin with age. International Journal of Dermatology, 23, 322–329.

    Article  PubMed  Google Scholar 

  • Levey, A. B. (1980). Measurement units in psychophysiology. In I. Martin & P. H. Venables (Eds.), Techniques in psychophysiology (pp. 597–628). New York: Wiley.

    Google Scholar 

  • Levinson, D. F., & Edelberg, R. (1985). Scoring criteria for response latency and habituation in electrodermal research: A critique. Psychophysiology, 22, 417–426.

    Article  PubMed  Google Scholar 

  • Lieblich, I., Kugelmass, S., & Ben-Shakhar, G. (1973). Psychophysiological baselines as a function of race and ethnic origin. Psychophysiology, 10, 426–430.

    Article  PubMed  Google Scholar 

  • Lim, C. L., Gordon, E., Rennie, C., Wright, J. J., Bahramali, H., Li, W. M., et al. (1999). Dynamics of SCR, EEG, and ERP activity in an odball paradigm with short interstimulus intervals. Psychophysiology, 36, 543–551.

    Article  PubMed  Google Scholar 

  • Lim, C. L., Rennie, C., Barry, J., Bahramali, H., Lazzaro, I., Manor, B., et al. (1997). Decomposing skin conductance into tonic and phasic components. International Journal of Psychophysiology, 25, 97–109.

    Article  PubMed  Google Scholar 

  • Lobstein, T., & Cort, J. (1978). The relationship between skin temperature and skin conductance activity: Indications of genetic and fitness determinants. Biological Psychology, 7, 139–143.

    Article  PubMed  Google Scholar 

  • Lockhart, R. A. (1972). Interrelations between amplitude, latency, rise time, and the Edelberg recovery measure of the galvanic skin response. Psychophysiology, 9, 437–442.

    Article  PubMed  Google Scholar 

  • Lowry, R. (1977). Active circuits for direct linear measurement of skin resistance and conductance. Psychophysiology, 14, 329–331.

    Article  PubMed  Google Scholar 

  • Lüer, G., & Neufeldt, B. (1967). Über Zeit- und Höhenmaße der galvanischen Hautreaktion. Psychologische Forschung, 30, 400–402.

    Article  PubMed  Google Scholar 

  • Lüer, G., & Neufeldt, B. (1968). Über den Zusammenhang zwischen Maßen der galvanischen Hautreaktion und Beurteilungen von Reizen durch Versuchspersonen. Zeitschrift für Experimentelle und Angewandte Psychologie, 15, 619–648.

    PubMed  Google Scholar 

  • Lykken, D. T. (1959a). Properties of electrodes used in electrodermal measurement. Journal of Comparative and Physiological Psychology, 52, 629–634.

    Article  PubMed  Google Scholar 

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

    Article  Google Scholar 

  • Lykken, D. T. (1982). Research with twins: The concept of emergenesis. Psychophysiology, 19, 361–373.

    Article  PubMed  Google Scholar 

  • Lykken, D. T., Iacono, W. G., Haroian, K., McGue, M., & Bouchard, T. J. (1988). Habituation of the skin conductance response to strong stimuli: A twin study. Psychophysiology, 25, 4–15.

    Article  PubMed  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., Miller, R. D., & Strahan, R. F. (1968). Some properties of skin conductance and potential. Psychophysiology, 5, 253–268.

    Article  PubMed  Google Scholar 

  • Lykken, D. T., & Rose, R. (1959). A rat-holder with electrodes for GSR measurement. The American Journal of Psychology, 72, 621–622.

    Article  PubMed  Google Scholar 

  • Lykken, D. T., Rose, R., Luther, B., & Maley, M. (1966). Correcting psychophysiological measures for individual differences in range. Psychological Bulletin, 66, 481–484.

    Article  PubMed  Google Scholar 

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

    Article  PubMed  Google Scholar 

  • Mahon, M. L., & Iacono, W. G. (1987). Another look at the relationship of electrodermal activity to electrode contact area. Psychophysiology, 24, 216–222.

    Article  PubMed  Google Scholar 

  • Malmo, R. B. (1965). Finger- sweat prints in the differentiation of low and high incentive. Psychophysiology, 1, 231–240.

    Article  PubMed  Google Scholar 

  • Malmstrom, E. J. (1968). The effect of prestimulus variability upon physiological reactivity scores. Psychophysiology, 5, 149–165.

    Article  PubMed  Google Scholar 

  • Maltzman, I., Gould, J., Barnett, O. J., Raskin, D. C., & Wolff, C. (1979). Habituation of the GSR and digital vasomotor components of the orienting reflex as a consequence of task instructions and sex differences. Physiological Psychology, 7, 213–220.

    Google Scholar 

  • Martin, I., & Rust, J. (1976). Habituation and the structure of the electrodermal system. Psychophysiology, 13, 554–562.

    Article  PubMed  Google Scholar 

  • Martínez-Selva, J. M., Gómez-Amor, J., Olmos, E., Navarro, N., & Román, F. (1987). Sex and menstrual cycle differences in the habituation and spontaneous recovery of the electrodermal orienting reaction. Personality and Individual Differences, 8, 211–217.

    Article  Google Scholar 

  • Maulsby, R. L., & Edelberg, R. (1960). The interrelationship between the galvanic skin response, basal resistance, and temperature. Journal of Comparative and Physiological Psychology, 53, 475–479.

    Article  PubMed  Google Scholar 

  • Mazurski, E. J., Bond, N. W., Siddle, D. A. T., & Lovibond, P. F. (1996). Conditioning with facial expressions of emotion: Effects of CS sex and age. Psychophysiology, 33, 416–425.

    Article  PubMed  Google Scholar 

  • McClendon, J. F., & Hemingway, A. (1930). The psychogalvanic reflex as related to the polarization-capacity of the skin. The American Journal of Psychology, 84, 77–83.

    Google Scholar 

  • McDowd, J. M., & Filion, D. L. (1992). Aging, selective attention, and inhibitory processes: A psychophysiological approach. Psychology and Aging, 7, 65–71.

    Article  PubMed  Google Scholar 

  • Mednick, S. A., & Schulsinger, F. (1968). Some premorbid characteristics related to breakdown in children with schizophrenic mothers. In S. Kety & D. Rosenthal (Eds.), The transmission of schizophrenia (pp. 267–291). Oxford: Pergamon.

    Google Scholar 

  • Miller, L. H., & Shmavonian, B. H. (1965). Replicability of two GSR indices as a function of stress and cognitive activity. Journal of Personality and Social Psychology, 2, 753–756.

    Article  PubMed  Google Scholar 

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

    Google Scholar 

  • Mitchell, D. A., & Venables, P. H. (1980). The relationship of EDA to electrode size. Psychophysiology, 17, 408–412.

    Article  PubMed  Google Scholar 

  • Mize, M. M., Vila-Coro, A. A., & Prager, T. C. (1989). The relationship between postnatal skin maturation and electrical skin impedance. Archives of Dermatology, 125, 647–650.

    Article  PubMed  Google Scholar 

  • Montagu, J. D. (1958). The psycho-galvanic reflex: A comparison of AC skin resistance and skin potential changes. Journal of Neurology, Neurosurgery, and Psychiatry, 21, 119–128.

    Article  PubMed  Google Scholar 

  • Montagu, J. D. (1973). The measurement of electrodermal activity: An instrument for recording log skin admittance. Biological Psychology, 1, 161–166.

    Article  PubMed  Google Scholar 

  • Montagu, J. D., & Coles, E. M. (1968). Mechanism and measurement of the galvanic skin response: An addendum. Psychological Bulletin, 69, 74–76.

    Article  Google Scholar 

  • Morimoto, T. (1978). Variations of sweating activity due to sex, age and race. In A. Jarrett (Ed.), The physiology and pathophysiology of the skin (The sweat glands, skin permeation, lymphatics, and the nails, Vol. 5, pp. 1655–1666). New York: Academic.

    Google Scholar 

  • Morkrid, L., & Qiao, Z.-G. (1988). Continuous estimation of parameters in skin electrical admittance from simultaneous measurements at two different frequencies. Medical & Biological Engineering & Computing, 26, 633–640.

    Article  Google Scholar 

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

    Google Scholar 

  • Muthny, F. A., Foerster, F., Hoeppner, V., Mueller, W., & Walschburger, P. (1983). Skin evaporative water loss (SE) and skin conductance (SC) under various psychophysiological conditions. Biological Psychology, 16, 241–253.

    Article  PubMed  Google Scholar 

  • Myrtek, M., & Foerster, F. (1986). The law of initial value: A rare exception. Biological Psychology, 22, 227–237.

    Article  PubMed  Google Scholar 

  • Myrtek, M., Foerster, F., & Wittmann, W. (1977). Das Ausgangswertproblem: Theoretische Überlegungen und empirische Untersuchungen. Zeitschrift für Experimentelle und Angewandte Psychologie, 24, 463–491.

    PubMed  Google Scholar 

  • Neufeld, R. W. J., & Davidson, P. O. (1974). Sex differences in stress response: A multivariate analysis. Journal of Abnormal Psychology, 83, 178–185.

    Article  PubMed  Google Scholar 

  • Neumann, E. (1968). Thermal changes in palmar skin resistance patterns. Psychophysiology, 5, 103–111.

    Article  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.

    Article  Google Scholar 

  • O’Connell, D. N., Tursky, B., & Evans, F. J. (1967). Normality of distribution of resting palmar skin potential. Psychophysiology, 4, 151–155.

    Article  PubMed  Google Scholar 

  • O’Gorman, J. G., & Horneman, C. (1979). Consistency of individual differences in non-specific electrodermal activity. Biological Psychology, 9, 13–21.

    Article  PubMed  Google Scholar 

  • Ödman, S. (1981). Potential and impedance variations following skin deformation. Medical & Biological Engineering & Computing, 19, 271–278.

    Article  Google Scholar 

  • Paintal, A. S. (1951). A comparison of the galvanic skin responses of normals and psychotics. Journal of Experimental Psychology, 41, 425–428.

    Article  PubMed  Google Scholar 

  • Pasquali, E., & Roveri, R. (1971). Measurement of the electrical skin resistance during skin drilling. Psychophysiology, 8, 236–238.

    Article  PubMed  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, 1797–1806.

    Article  PubMed  Google Scholar 

  • Plouffe, L., & Stelmack, R. M. (1984). The electrodermal orienting response and memory: An analysis of age differences in picture recall. Psychophysiology, 21, 191–198.

    Article  PubMed  Google Scholar 

  • Plutchik, R., & Hirsch, H. R. (1963). Skin impedance and phase angle as a function of frequency and current. Science, 141, 927–928.

    Article  PubMed  Google Scholar 

  • Poh, M. Z., Loddenkemper, T., Swenson, N. C., Goyal, S., Madsen, J. R., & Picard, R. W. (2010b). Continuous monitoring of electrodermal activity during epileptic seizures using a wearable sensor. 32nd Annual International Conference of the IEEE EMBS, Buenos Aires, Argentina (pp. 4415–4418) IEEE EBMS: Piscataway, NJ.

    Google Scholar 

  • Poh, M. Z., Swenson, N. C., & Picard, R. W. (2010). A wearable sensor for unobtrusive, long-term assessment of electrodermal activity. IEEE Transactions on Biomedical Engineering, 57, 1243–1252.

    Article  PubMed  Google Scholar 

  • Pollack, S. V. (1985). The aging skin. The Journal of the Florida Medical Association, 72, 245–248.

    PubMed  Google Scholar 

  • Potts, R. O., Buras, E. M., & Chrismas, D. A. (1984). Changes with age in the moisture content of human skin. The Journal of Investigative Dermatology, 82, 97–100.

    Article  PubMed  Google Scholar 

  • Prokasy, W. F., & Kumpfer, K. L. (1973). Classical conditioning. In W. F. Prokasy & D. C. Raskin (Eds.), Electrodermal activity in psychological research (pp. 157–202). New York: Academic.

    Google Scholar 

  • Purohit, A. P. (1966). Personality variables, sex differences, GSR responsiveness and GSR conditioning. Journal of Experimental Research in Personality, 1, 165–179.

    Google Scholar 

  • Quiao, Z. G., Morkrid, L., & Grimnes, S. (1987). Simultaneous measurement of electrical admittance, blood flow and temperature at the same skin site with a specially designed probe. Medical & Biological Engineering & Computing, 25, 299–304.

    Article  Google Scholar 

  • Rachman, S. (1960). Reliability of galvanic skin response measures. Psychological Reports, 6, 326.

    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.

    Article  PubMed  Google Scholar 

  • Raine, A., & Venables, P. H. (1984). Electrodermal nonresponding, antisocial behavior, and schizoid tendencies in adolescents. Psychophysiology, 21, 424–433.

    Article  PubMed  Google Scholar 

  • Rickles, W. H., & Day, J. L. (1968). Electrodermal activity in non-palmar skin sites. Psychophysiology, 4, 421–435.

    Article  PubMed  Google Scholar 

  • Román, F., García-Sánchez, F. A., Martínez-Selva, J. M., Gómez-Amor, J., & Carrillo, E. E. (1989). Sex differences and bilateral electrodermal activity: A replication. The Pavlovian Journal of Biological Science, 24, 150–155.

    PubMed  Google Scholar 

  • Rutenfranz, J. (1955). Zur Frage einer Tagesrhythmik des elektrischen Hautwiderstandes beim Menschen. Internationale Zeitschrift für angewandte Physiologie einschließlich Arbeitsphysiologie, 16, 152–172.

    Google Scholar 

  • Rutenfranz, J. (1958). Der Widerstand der Haut gegenüber schwachen elektrischen Strömen. Der Hautarzt, 9, 289–299.

    Google Scholar 

  • Rutenfranz, J., & Wenzel, H. G. (1958). Über quantitative Zusammenhänge zwischen Wasserabgabe, Wechselstromwiderstand und Kapazität der Haut bei körperlicher Arbeit und unter verschiedenen Raumtemperaturen. Internationale Zeitschrift für angewandte Physiologie einschließlich Arbeitsphysiologie, 17, 155–176.

    Google Scholar 

  • Sagberg, F. (1980). Dependence of EDR recovery times and other electrodermal measures on scale of measurement: A methodological clarification. Psychophysiology, 17, 506–509.

    Article  PubMed  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 

  • Sato, F., Owen, M., Matthes, R., Sato, K., & Gisolfi, C. V. (1990). Functional and morphological changes in the eccrine sweat gland with heat acclimation. Journal of Applied Physiology, 69, 232–236.

    PubMed  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, 241–246.

    Article  PubMed  Google Scholar 

  • Schaefer, F. (1993). A new approach to circumventing the conductance-resistance choice: Phase angle between alternating-current and -voltage. In J. C. Roy, W. Boucsein, D. C. Fowles, & J. H. Gruzelier (Eds.), Progress in electrodermal research (pp. 43–48). London: Plenum.

    Chapter  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.

    Article  PubMed  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 

  • Schneider, R. L. (1987). A mathematical model of human skin conductance. Psychophysiology, 24, 610.

    Google Scholar 

  • Schneider, R. E., & Fowles, D. C. (1978). A convenient, non-hydrating electrolyte medium for the measurement of electrodermal activity. Psychophysiology, 15, 483–486.

    Article  PubMed  Google Scholar 

  • Schneider, R., Schmidt, S., Binder, M., Schaefer, F., & Walach, H. (2003). Respiration-related artifacts in EDA recordings: Introducing a standardized method to overcome multiple interpretations. Psychological Reports, 93, 907–920.

    PubMed  Google Scholar 

  • Schönpflug, W., Deusinger, I. M., & Nitsch, F. (1966). Höhen – und Zeitmaße der psychogalvanischen Reaktion. Psychologische Forschung, 29, 1–21.

    Article  PubMed  Google Scholar 

  • Schulter, G., & Papousek, I. (1992). Bilateral electrodermal activity: Reliability, laterality and individual differences. International Journal of Psychophysiology, 13, 199–213.

    Article  PubMed  Google Scholar 

  • Schwan, H. P. (1963). Determination of biological impedances. In W. L. Nastuk (Ed.), Physical techniques in biological research (Vol. 6, pp. 323–407). New York: Academic.

    Google Scholar 

  • Shackel, B. (1959). Skin-drilling: A method of diminishing galvanic skin potentials. The American Journal of Psychology, 72, 14–21.

    Article  Google Scholar 

  • Shapiro, D., & Leiderman, P. H. (1954). Studies on the galvanic skin potential level: Some statistical properties. Journal of Psychosomatic Research, 7, 269–275.

    Article  Google Scholar 

  • Shmavonian, B. M., & Busse, E. W. (1963). Psychophysiologic techniques in the study of the aged. In R. Williams, C. Tibbits, & W. Donahue (Eds.), Processes of aging (pp. 160–183). New York: Atherton.

    Google Scholar 

  • Shmavonian, B. M., Miller, L. H., & Cohen, S. I. (1968). Differences among age and sex groups in electrodermal conditioning. Psychophysiology, 5, 119–131.

    Article  PubMed  Google Scholar 

  • Shmavonian, B. M., Yarmat, A. J., & Cohen, S. I. (1965). Relationship between the autonomic nervous system and central nervous system in age differences in behavior. In A. T. Welford & J. E. Birren (Eds.), Aging and the nervous system (pp. 235–258). Springfield: Thomas.

    Google Scholar 

  • Silver, A., Montagna, W., & Karacan, I. (1965). The effect of age on human eccrine sweating. In W. Montagna (Ed.), Advances in biology of skin (Vol. 6, pp. 129–150). Oxford: Pergamon.

    Google Scholar 

  • Silverman, A. J., Cohen, S. I., & Shmavonian, B. M. (1958). Psychophysiologic response specificity in the elderly. Journal of Gerontology, 5, 443.

    Google Scholar 

  • Silverman, A. J., Cohen, S. I., & Shmavonian, B. M. (1959). Investigation of psychophysiological relationships with skin resistance measures. Journal of Psychosomatic Research, 4, 65–87.

    Article  PubMed  Google Scholar 

  • Silverman, J. J., & Powell, V. E. (1944). Studies on palmar sweating: I. A technique for the study of palmar sweating. American Journal of the Medical Sciences, 208, 297–299.

    Article  Google Scholar 

  • Simon, W. R., & Homoth, R. W. G. (1978). An automatic voltage suppressor for the measurement of electrodermal activity. Psychophysiology, 15, 502–505.

    Article  PubMed  Google Scholar 

  • Simpson, A., & Turpin, G. (1983). A device for ambulatory skin conductance monitoring. Psychophysiology, 20, 225–229.

    Article  PubMed  Google Scholar 

  • Sokolov, E. N. (1963a). Higher nervous functions: The orienting reflex. Annual Review of Physiology, 25, 545–580.

    Article  PubMed  Google Scholar 

  • Sokolov, E. N. (1963b). Perception and the conditioned reflex. Oxford: Pergamon.

    Google Scholar 

  • Spinks, J. A., Dow, R., & Chiu, L. W. (1983). A microcomputer package for real-time skin conductance response analysis. Behavior Research Methods & Instrumentation, 15, 591–593.

    Article  Google Scholar 

  • Steigleder, G. K. (1983). Dermatologie und Venerologie. Stuttgart: Thieme.

    Google Scholar 

  • Stemmler, G. (1984). Psychophysiologische Emotionsmuster. Frankfurt: Peter Lang.

    Google Scholar 

  • Stemmler, G. (1987). Standardization within subjects: A critique of Ben-Shakhar’s conclusions. Psychophysiology, 24, 243–246.

    Article  PubMed  Google Scholar 

  • Stephens, W. G. S. (1963). The current-voltage relationship in human skin. Medical, Electronics and Biological Engineering, 1, 389–399.

    Article  Google Scholar 

  • Stern, R. M., & Anschel, C. (1968). Deep inspirations as stimuli for responses of the autonomic nervous system. Psychophysiology, 5, 132–141.

    Article  PubMed  Google Scholar 

  • Stern, J. A., & Walrath, L. C. (1977). Orienting responses and conditioning of electrodermal responses. Psychophysiology, 14, 334–342.

    Article  PubMed  Google Scholar 

  • Sternbach, R. A., & Tursky, B. (1965). Ethnic differences among housewives in psychophysical and skin potential responses to electric shock. Psychophysiology, 1, 241–246.

    Article  PubMed  Google Scholar 

  • Surwillo, W. W. (1965). Level of skin potential in healthy males and the influence of age. Journal of Gerontology, 20, 519–521.

    PubMed  Google Scholar 

  • Surwillo, W. W. (1967). The influence of some psychological factors on latency of the galvanic skin reflex. Psychophysiology, 4, 223–228.

    Article  PubMed  Google Scholar 

  • Surwillo, W. W. (1969). Statistical distribution of volar skin potential level in attention and the effects of age. Psychophysiology, 6, 13–16.

    Article  PubMed  Google Scholar 

  • Surwillo, W. W., & Quilter, R. E. (1965). The relation of frequency of spontaneous skin potential responses to vigilance and to age. Psychophysiology, 1, 272–276.

    Article  PubMed  Google Scholar 

  • Sutarman, & Thomson, M. L. A. (1952). A new technique for enumerating active sweat glands in man. Journal of Physiology, 117 , 51–52.

    Google Scholar 

  • Swartzman, L. C., Edelberg, R., & Kemmann, E. (1990). The menopausal hot flush: Symptom reports and concomitant physiological changes. Journal of Behavioral Medicine, 13, 15–30.

    Article  PubMed  Google Scholar 

  • Tassinary, L. G., Geen, T. R., Cacioppo, J. T., & Edelberg, R. (1990). Issues in biometrics: Offset potentials and the electrical stability of Ag/AgCl electrodes. Psychophysiology, 27, 236–242.

    Article  PubMed  Google Scholar 

  • Thetford, P. E., Klemme, M. E., & Spohn, H. E. (1968). Skin potential, heart rate, and the span of immediate memory. Psychophysiology, 5, 166–177.

    Article  PubMed  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 

  • Thom, E. (1988). Die Hamburger EDA-Auswertung. In W. Boucsein (Ed.), Elektrodermale Aktiviät. Grundlagen, Methoden und Anwendungen (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 

  • Thompson, M. L. (1954). A comparison between the number and distribution of functioning eccrine sweat glands in Europeans and Africans. The Journal of Physiology, 123, 225–233.

    Google Scholar 

  • Toyokura, M. (1999). Waveform variation and size of sympathetic skin response: Regional difference between the sole and palm recordings. Clinical Neurophysiology, 110, 765–771.

    Article  PubMed  Google Scholar 

  • Toyokura, M. (2006). Sympathetic skin responses: The influence of electrical stimulus intensity and habituation on the waveform. Clinical Autonomic Research, 16, 130–135.

    Article  PubMed  Google Scholar 

  • Traxel, W. (1957). Über das Zeitmaß der psychogalvanischen Reaktion. Zeitschrift für Psychologie, 161, 282–291.

    PubMed  Google Scholar 

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

    Google Scholar 

  • Tronstad, C., Johnsen, G. K., Grimnes, S., & Martinsen, O. G. (2010). A study on electrode gels for skin conductance measurements. Physiological Measurement, 31, 1395–1410.

    Article  PubMed  Google Scholar 

  • Turpin, G., & Clements, K. (1993). Electrodermal activity and psychopathology: The development of the palmar sweat index (PSI) as an applied measure for use in clinical settings. In J.-C. Roy, W. Boucsein, D. C. Fowles, & J. H. Gruzelier (Eds.), Progress in electrodermal research: From physiology to psychology (pp. 49–60). London: Plenum.

    Google Scholar 

  • Turpin, G., Shine, P., & Lader, H. (1983). Ambulatory electrodermal monitoring effects of ambient temperature, general activity, electrolyte media, and length of recording. Psychophysiology, 20, 219–224.

    Article  PubMed  Google Scholar 

  • Uncini, A., Pullman, S. L., Lovelace, R. E., & Gambi, D. (1988). The sympathetic skin response: Normal values, elucidation of afferent components and application limits. Journal of the Neurological Sciences, 87, 299–306.

    Article  PubMed  Google Scholar 

  • Van Boxtel, A. (1977). Skin resistance during square-wave electrical pulses of 1 to 10 mA. Medical & Biological Engineering & Computing, 15, 679–687.

    Article  Google Scholar 

  • Venables, P. H. (1955). The relationships between P.G.R. scores and temperature and humidity. Quarterly Journal of Experimental Psychology, 7, 12–18.

    Article  Google Scholar 

  • Venables, P. H. (1978). Psychophysiology and psychometrics. Psychophysiology, 15, 302–314.

    Article  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 

  • Venables, P. H., & Fletcher, R. P. (1981). The status of skin conductance recovery time: An examination of the Bundy effect. Psychophysiology, 18, 10–16.

    Article  PubMed  Google Scholar 

  • Venables, P. H., Gartshore, S. A., & O’Riordan, P. W. (1980). The function of skin conductance response recovery and rise time. Biological Psychology, 10, 1–6.

    Article  PubMed  Google Scholar 

  • Venables, P. H., & Martin, I. (1967a). Skin resistance and skin potential. In P. H. Venables & I. Martin (Eds.), A manual of psychophysiological methods (pp. 53–102). Amsterdam: North Holland.

    Google Scholar 

  • Venables, P. H., & Martin, I. (1967b). The relation of palmar sweat gland activity to level of skin potential and conductance. Psychophysiology, 3, 302–311.

    Article  PubMed  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, 87–101.

    Article  PubMed  Google Scholar 

  • Venables, P. H., & Sayer, E. (1963). On the measurement of the level of skin potential. British Journal of Psychology, 54, 251–260.

    Article  PubMed  Google Scholar 

  • Waid, W. M. (1974). Degree of goal-orientation, level of cognitive activity and electrodermal recovery rate. Perceptual and Motor Skills, 38, 103–109.

    Article  PubMed  Google Scholar 

  • Walschburger, P. (1975). Zur Standardisierung und Interpretation elektrodermaler Messwerte in psychologischen Experimenten. Zeitschrift für Experimentelle und Angewandte Psychologie, 22, 514–533.

    PubMed  Google Scholar 

  • Walschburger, P. (1976). Zur Beschreibung von Aktivierungsprozessen: Eine Methodenstudie zur psychophysiologischen Diagnostik. Freiburg: Unpublished Doctoral Dissertation, Albert-Ludwig-Universität.

    Google Scholar 

  • Ward, N. G., & Doerr, H. O. (1986). Skin conductance: A potentially sensitive and specific marker for depression. The Journal of Nervous and Mental Disease, 174, 553–559.

    Article  PubMed  Google Scholar 

  • Ward, N. G., Doerr, H. O., & Storrie, M. C. (1983). Skin conductance: A potentially sensitive test for depression. Psychiatry Research, 10, 295–302.

    Article  PubMed  Google Scholar 

  • Waters, W. F., Koresko, R. L., Rossie, G. V., & Hackley, S. A. (1979). Short-, medium-, and long-term relationships among meteorological and electrodermal variables. Psychophysiology, 16, 445–451.

    Article  PubMed  Google Scholar 

  • Weigand, D. A., Haywood, C., & Gaylor, J. R. (1974). Cell layers and density of Negro and Caucasian stratum corneum. The Journal of Investigative Dermatology, 62, 563–568.

    Article  PubMed  Google Scholar 

  • Weisz, J., & Czigler, I. (2006). Age and novelty: Event-related brain potentials and autonomic activity. Psychophysiology, 43, 261–271.

    Article  PubMed  Google Scholar 

  • Wenger, M. A., & Cullen, T. D. (1962). Some problems in psychophysiological research: III. The effects of uncontrolled variables. In R. Roessler & N. S. Greenfield (Eds.), Psychophysiological correlates of psychological disorder (pp. 106–114). Madison: University of Wisconsin Press.

    Google Scholar 

  • Westerink, J., Ouwerkerk, M., de Vries, G.-J., de Waele, S., van den Eerenbeemd, J., & van Boven, M. (2009, 10–12 Sept). Emotion measurement platform for daily life situations. International Conference on Affective Computing and Intelligent Interaction (ACII), Amsterdam, The Netherlands.

    Google Scholar 

  • Wieland, B. A., & Mefferd, R. B. (1970). Systematic changes in levels of physiological activity during a four-month period. Psychophysiology, 6, 669–689.

    Article  PubMed  Google Scholar 

  • Wilcott, R. C. (1958). Correlation of skin resistance and potential. Journal of Comparative and Physiological Psychology, 51, 691–696.

    Article  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.

    Article  PubMed  Google Scholar 

  • Wilcott, R. C. (1964). The partial independence of skin potential and skin resistance from sweating. Psychophysiology, 1, 55–66.

    Article  PubMed  Google Scholar 

  • Wilcott, R. C., & Hammond, L. J. (1965). On the constancy-current error in skin resistance measurement. Psychophysiology, 2, 39–41.

    Article  PubMed  Google Scholar 

  • Wilder, J. (1931). Das”Ausgangswert-Gesetz” – ein unbeachtetes biologisches Gesetz; seine Bedeutung für Forschung und Praxis. Klinische Wochenschrift, 41, 1889–1893.

    Article  Google Scholar 

  • Wilhelm, F. H., & Roth, W. T. (1996). Ambulatory assessment of clinical anxiety. In J. Fahrenberg & M. Myrtek (Eds.), Ambulatory assessment: Computer-assisted psychological and psychophysiological methods in monitoring and field studies (pp. 317–345). Göttingen: Hogrefe.

    Google Scholar 

  • Wilhelm, F. H., Roth, W. T., & Sackner, M. (2003). The LifeShirt: An advanced system for ambulatory measurement of respiratory and cardiac function. Behavior Modification, 27, 671–691.

    Article  PubMed  Google Scholar 

  • Williams, L. M., Brammer, M. J., Skerrett, D., Lagopolous, J., Rennie, C., Kozek, K., et al. (2000). The neural correlates of orienting: An integration of fMRI and skin conductance orienting. Neuroreport, 11, 3011–3015.

    Article  PubMed  Google Scholar 

  • Williams, L. M., Brown, K. J., Das, P., Boucsein, W., Sokolov, E. N., Brammer, M. J., et al. (2004). The dynamics of cortico-amygdala and automatic activity over the experimental time course of fear perception. Brain Research. Cognitive Brain Research, 21, 114–123.

    Article  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.

    Article  PubMed  Google Scholar 

  • Wyatt, R., & Tursky, B. (1969). Skin potential levels in right- and left-handed males. Psychophysiology, 6, 133–137.

    Article  PubMed  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, Y., & Yamamoto, T. (1978). Technical note: Dispersion and correlation of the parameters for skin impedance. Medical & Biological Engineering & Computing, 16, 592–594.

    Article  Google Scholar 

  • Yamamoto, Y., & Yamamoto, T. (1979). Technical note: Dynamic system for the measurement of electrical skin impedance. Medical & Biological Engineering & Computing, 17, 135–137.

    Article  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.

    Article  Google Scholar 

  • Yokota, T., & Fujimori, B. (1962). Impedance change of the skin during the galvanic skin reflex. The Japanese Journal of Physiology, 12, 200–209.

    Article  PubMed  Google Scholar 

  • Zahn, T. P. (1976). On the bimodality of the distribution of electrodermal orienting responses in schizophrenic patients. The Journal of Nervous and Mental Disease, 162, 195–199.

    Article  PubMed  Google Scholar 

  • Zelinski, E. M., Walsh, D. A., & Thompson, L. W. (1978). Orienting task effects on EDR and free recall in three age groups. Journal of Gerontology, 33, 239–245.

    PubMed  Google Scholar 

  • Zimmer, H. (2000). Frequenz und mittlere Amplitude spontaner elektrodermaler Fluktuationen sind keine austauschbaren Indikatoren psychischer Prozesse. Zeitschrift für Experimentelle Psychologie, 47, 129–143.

    PubMed  Google Scholar 

  • Zipp, P. (1983). Impedance controlled skin drilling. Medical & Biological Engineering & Computing, 21, 382–384.

    Article  Google Scholar 

  • Zipp, P., & Faber, S. (1979). Rückwirkungsarme Ableitung bioelektrischer Signale bei arbeitswissenschaftlichen Langzeituntersuchungen am Arbeitsplatz. European Journal of Applied Physiology and Occupational Physiology, 42, 105–116.

    Article  PubMed  Google Scholar 

  • Zipp, P., Hennemann, K., Grunwald, R., & Rohmert, W. (1980). Bewertung von Kontaktvermittlern für Bioelektroden bei Langzeituntersuchungen. European Journal of Applied Physiology and Occupational Physiology, 45, 131–145.

    Article  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). Methods of Electrodermal Recording. In: Electrodermal Activity. Springer, Boston, MA. https://doi.org/10.1007/978-1-4614-1126-0_2

Download citation

Publish with us

Policies and ethics