Skip to main content

Applications of Electrodermal Recording

  • Chapter
  • First Online:
Electrodermal Activity

Abstract

The third part of this book is dedicated to various applications of EDA recording. The aim is to provide a theoretical framework for the use of the different EDA parameters described in Chap. 2 as psychophysiological indicators in the appropriate fields. Since there are thousands of articles reporting EDA results (Sect. 1.1.3), their comprehensive description would go far beyond the limits of the present book. Instead, the focus will be on giving more detailed information especially for studies which enlighten either methodological issues or provide support for interpretation of results in the light of psychophysiological theories related to EDA.

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.

    For summaries of older results, see Prokasy and Raskin (1973) and Edelberg (1972a). More recent reviews will be referred to within the appropriate sections.

  2. 2.

    During the 1970s, in the wake of the cognitive “paradigm shift” (Sect. 3.1.2.1 “Cognitive vs. Noncognitive View of EDR Conditioning”), comparator theories of OR have been partly rephrased in terms of resource allocation in a system with limited capacity (Sect. 3.1.2.1 “Components of EDR Conditioning”). A brief description of Öhman’s (1979) and his own model is provided by Hugdahl (1995, pp. 135–138). The possible role of the SC-OR being a potential indicator of resource allocation will be discussed in Sects. 3.1.3.1 and 3.1.3.4. For more information on SC-OR and resource allocation see, for example, Siddle (1991), Siddle, Lipp, and Dall (1996) and Filion, Dawson, Schell, and Hazlett (1991), described in Sect. 3.1.3.2; critical results were reported by Niepel (2001).

  3. 3.

    The Groves-Thompson theory had been developed based on research on spinal cats. Since the proposed sensitization process has not been unambiguously demonstrated in studies with humans, such a conclusion is barely justified (see Sect. 3.1.1.3).

  4. 4.

    Below-zero habituation represents a continuation of stimulus presentation after reaching an individual habituation criterion, e.g., two successive “zero-responses” (Sect. 3.1.1.3).

  5. 5.

    In steps of 10 dB, each stimulus applied five times in a balanced design, SR recording with 50 μA constant current using 2 cm2 Ag electrodes from two fingers of the right hand, transformed to SC.

  6. 6.

    In steps of 20 dB; SCR recorded with standard electrodes and voltage, however unipolar thenar against a neutral forearm side and with K–Y gel (Experiment 3).

  7. 7.

    Recorded AC coupled (Sect. 2.1.3) with a 5 s time constant volar from the left hand’s fingers.

  8. 8.

    As already described at the end of Sect. 2.5.1.1, Toyokura (2006) obtained a directly proportional relationship between the intensity of an electrical stimulus applied to the median nerve at the wrist and SPR amp., being either positively or negatively directed (Sect. 2.3.1.2 “Amplitudes of Endosomatic Responses”) in 40 male participants. This relationship persisted even after habituation. Furthermore, OR reinstatement was clearly observable if the electric stimulation was switched to magnetic stimulation.

  9. 9.

    Rotenberg and Vedenyapin (1985) in a study with 15 participants being presented a series of tones, a subset of whom had to react to the tone, found some evidence of SPR amp. to the tones being more dependent on decision making than on motor preparation.

  10. 10.

    Recorded as SRR palmar, 0.5–5.0 s after stimulus onset and subjected to a logarithmic SCR transformation.

  11. 11.

    Measured with stainless steel 2 × 3 cm electrodes from the volar surface of the left hand’s fingers with Biocom Inc. Biogel as contact medium; 1–5 s following stimulus onset was used as time window.

  12. 12.

    Stimuli were one-word (name of an occupation) or two-word stimuli (occupation name combined with name of a hobby), both kinds mutually used as standard and as test stimuli. The participants had to recall as many words as possible. Receiver operating characteristic (ROC) curves were generated by comparing the distributions of standardized responses to test stimuli vs. standard stimuli.

  13. 13.

    A name of a hobby instead of a certain number, as used in an earlier experiment (Ben-Shakhar & Lieblich 1982), had a clear common component shared with the standard stimuli (which were also numbers).

  14. 14.

    Recorded with 12.55 mm diameter Ag/AgCl electrodes filled with K–Y jelly from thenar/hypothenar sites of the nondominant hand, using 1.0 V DC and a Wheatstone bridge. Two separate channels were used for SCL and SCR to enable a full range of 5 μS for the response channel (Sects. 2.1.3 and 2.2.4.1). SCR amp. were scored as changes in SC from the prestimulus level to the response peak within a 3–7 s window after the presentation of words.

  15. 15.

    Recorded bilaterally from the hypothenar eminence with standard methodology, using Beckman miniature electrodes, logarithmically transformed.

  16. 16.

    Bundy’s “X” (Sect. 2.5.2.5, (2.22)) and SCR rec.t/2 yielded significant correlations for 3 of the 10 participants.

  17. 17.

    During the last 2 decades, SC-OR has been frequently used to determine the OR component of a so-called prepulse, which has been applied to modify the human startle eyeblink reflex (Filion, Dawson, & Schell, 1998). Such a stimulus either terminates before, or continues up to, and sometimes beyond the startle stimulus (Wynn, Dawson, & Schell, 2000). It was hypothesized that a prolonged prepulse may elicit a generalized OR (Sect. 3.1.1).

  18. 18.

    Recorded between an active palmar and an inactive wrist site with Ag electrodes being 2.6 cm in diameter.

  19. 19.

    Between 40 and 120 dB in steps of 20 dB, however with independent groups of 25 participants each. Endosomatic EDA was recorded from left thenar against pretreated forearm sites, while exosomatic measures were taken as SR from the right hand palmar vs. dorsal with 40 μA, being transformed to SC values, using Beckman Ag/AgCl electrodes with NaCl paste for both measures.

  20. 20.

    Although not aiming at the OR/DR distinction, Toyokura (2006) demonstrated in 40 healthy participants that a positive SPR component was more likely to be observed in high-intensity compared to low-intensity median nerve electrical stimulation. In the second out of four consecutive sessions, the predominant SPR component switched from positivity to negativity, which the author interpreted as an effect of habituation. However, the stimulation sequence was not suitable for investigating habituation as described in Sect. 3.1.1.3. A sort of OR reinstatement was observed when introducing a magnetic stimulation to the neck, which also showed no habituation of the SPR. SP (labeled SSR, see Sect. 3.5.4) was recorded from the right midpalm against the forearm.

  21. 21.

    In 15 dB steps, 30 ms stimulus rise time and 2 s duration. EDA was recorded with standard methodology, however using K–Y gel.

  22. 22.

    Presenting 1 kHz tones between 80 and 120 dB intensity, in steps of 10 dB.

  23. 23.

    Recorded with 32 Hz and 8 μA via 1 cm2 Ag electrodes, using paper soaked with 0.5% NaCl as an electrolyte, dorsal from the first and third finger, and palmar from the second and fourth finger of the nondominant hand.

  24. 24.

    Recorded with standard methodology from the middle phalanges of the index and middle finger of the left hand. SCR amp. with latencies between 1 and 5 s after stimulus onset and an amplitude exceeding 0.02 μS were evaluated and log-transformed (Sect. 2.3.3.3).

  25. 25.

    Providing such an amplitude criterion is also of fundamental significance for investigations into the so-called “below-zero” habituation (Thompson & Spencer, 1966), which also requires taking into account the sensitivity of the physiological system in question (Stephenson & Siddle, 1976). However, there is only weak evidence from studies with humans for an influence of the duration of “below-zero” habituation on the spontaneous recovery of an OR after being habituated (i.e., OR reinstatement; Siddle, Remington, Kuiack, & Haines 1983).

  26. 26.

    Hölzl, Wilhelm, Lutzenberger, and Schandry (1975, Fig. 12) in their study described in Sect. 2.6.5 observed three types of SCR courses in their 28 participants. Most of them showed an exponential decrease over trials, a few “sensitizers” exhibited a slight increase, and some other “initial sensitizers” showed an increase followed by a considerable decrease.

  27. 27.

    This result gives rise to the hitherto unanswered question of the adequate unit of measurement in EDR habituation studies (Sect. 2.6.5). However, the habituation effects in this study may have been obscured by the fact that a counterbalanced presentation of different stimulus intensities had been applied. This is different from standard habituation series, which may have contributed to the marked dishabituation effects observed with the high-intensity stimuli.

  28. 28.

    Many authors combine two or more trials forming a trial block when reporting habituation courses.

  29. 29.

    For example, if significantly smaller values for a as well as for b are observed in a group with older participants as compared to a younger group, and if the age effects on the ordinate value a and on the gradient b are independent of each other as well as of the correlation between a and b, the correction of the gradient b according to (3.1), taking into account the ordinate value a, will lead to a reduction of the “independent” age effect on gradient b. In general, it must be regarded as difficult to determine the relative influence of experimental conditions or somatic variables like age and gender on values of a and b. Therefore, if those conditions have a significant influence on a, corrections according to (3.1) should be avoided.

  30. 30.

    SCR recorded with standard methodology during the presentation of 1 kHz tones (70 or 90 dB). The signs from Siddle and Heron’s (1976) Table 2 were inverted to make the directions of correlations comparable to each other.

  31. 31.

     KHz, 65 dB, 5 s duration tone. EDA was recorded as SR with 1 cm2 Ag/AgCl electrodes and NaCl paste, using 11 μA current, transformed into log SC units.

  32. 32.

    Three successive SCRs recorded with standard methodology, showing SCR amp. below 0.02 μS.

  33. 33.

    The author thanks Professor John J. Furedy for his contributions to this section, especially to its historical part.

  34. 34.

    The problems with this design, from the perspective of the cognitive contingency paradigm, are discussed in detail by Grings, Givens, and Carey (1979), Rescorla (1967), and Seligman (1969). However, from a consideration of the first fact regarding EDR conditioning, noted below, these methodological problems do not apply to the within-subjects, differential conditioning design. The practical advantage of this design is its greater sensitivity, which is important for the EDR conditioning phenomenon. If a between-subjects design is used, the conditioning/control difference requires about 20 or more participants just to obtain the basic conditioning phenomenon at the 0.05 level of significance within humans, partly because the aversive UCS (shock or noise) has to be kept at relatively mild levels for ethical reasons. This practical advantage is much less important in the older human eyelid conditioning preparation (for an account, see Kimble, 1961), which is more robust and allows observation of gradually increasing acquisition functions over as many as 100 trials; this preparation has mostly used the between-subjects design.

  35. 35.

    For ethical reasons, it came into use in the recent practice of aversive EDR conditioners to have participants set their own level of UCS intensity, rather than the older practice of the experimenter setting UCS intensity for all participants. This practice reduces the robustness of the ED-CR phenomenon, as well as making it more difficult to vary UCS intensity itself, which used to be one of the basic independent variables of human eyelid conditioning (Spence, Haggard, & Ross, 1958), seldom if ever appears in current human EDR conditioning papers.

  36. 36.

    Electrodermal recording has been described as being taken from the left index finger and the right palm with a Fels dermohmeter.

  37. 37.

    Formerly abbreviated as FAR and SAR, respectively.

  38. 38.

    This argument may be questioned, based on findings obtained by Furedy and Scull (1971). They observed what they called a “relative refractory period” of at least 20 s for an EDR. However, in the present author’s view, this term could be misleading, since its physiological meaning refers to nerve discharge in time windows of millisecond duration. Instead, the observed phenomena might be discussed in terms of relationships between EDR amp. following each other in short sequence (Sects. 2.3.1.5 and 2.5.2.5). A decrease of SCR amp. with shortening the ISI of average 20 s length by 50% was recently observed by Breska, Maoz, and Ben-Shakhar (2011), in their study mentioned in Footnote 58, Sect. 2.3.1.5. Since the ISIs used in classical conditioning are even much shorter, the logical consequence would be that the larger the FIR, the smaller the following (by about 4 s) SIR, a factor which would contribute a negative correlation between FIR and SIR. On the other hand, in terms of trial-to-trial variation, one would expect a positive correlation between FIR and SIR. Accordingly, the sum of these two opposing factors may produce a zero correlation which does not necessarily indicate real independence between FIR and SIR. The eyelid conditioning case is different, as in this paradigm a clear independence between FIR (called alpha response) and SIR (called and scored as the true CR) is demonstrated inasmuch as the former decreases as a function of repeated CS–UCS trials (habituation), whereas the latter increases (CR acquisition).

  39. 39.

    A very strong version of this “reinstated” OR was used (e.g., Stewart et al., 1961) to reject the short-interval ISI preparation as a demonstration of EDR conditioning, a preparation that required CS-alone test trials during acquisition to assess conditioning. The argument was that the EDRs to such CS-alone test trials were not CRs, but were merely “reinstated” ORs to stimulus change, rather than to classical EDR conditioning. Now it is true that marked stimulus change does increase the EDR, as in a crossmodal change following 15 repetitions (Furedy, 1968), and this could be reasonably referred to as the phenomenon of OR reinstatement. However, this change effect requires far greater quantitative and qualitative changes from repetition to change than are involved in CS-alone test trials during acquisition (CS–UCS pairings), so that the empirical evidence that supports the reinstated OR interpretation on UCS-omission trials is quite weak (Furedy & Poulos, 1977).

  40. 40.

    EDRs were recorded as SRRs using standard methodology, however with K–Y gel.

  41. 41.

    The role of awareness in classical electrodermal conditioning has also been extensively discussed by Dawson and Schell (1985, pp. 107 f.). They compared the technique used in various studies of their own group and by other researchers, embedding CS–UCS pairings within a “masking task,” with the method of CNS ablation studies, since the participant’s attention and concern is effectively directed away from the learning task. During the 1970s, the Dawson group published seven separate experiments involving more than 300 participants using the same masking task (reporting colors of lights and position of tones). They all showed that CS–UCS pairings were not sufficient to establish differential autonomic conditioning unless participants were aware of the contingency, the most suitable measure of which was a short recognition post-conditioning questionnaire. Their conclusion was that controlled cognitive processes are necessary for human autonomic discrimination in classical conditioning (see also Sect. 3.1.3.2).

  42. 42.

    There is also the possibility of a moderate heritability of electrodermal differential fear conditioning (Sect. 2.4.3.3).

  43. 43.

    Recorded from the fingers as SR with 2 cm2 Ag electrodes, NaCl paste, and 45 μA current strength, changed to SC values and square-root transformed. Two-second white noise of 100 dB served as UCS, while the CS was 80, 90, or 98 dB, varying between 2 and 10 s, in 2 s increments.

  44. 44.

    Recorded from the fingers as SR with 2.5 cm2 Ag electrodes, using 45 μA current, changed to SC and square-root transformed.

  45. 45.

    Recorded as SR from the medial phalanges using standard methodology. SRRs exceeding an amplitude criterion of 0.05 kΩ were transformed to SCRs and additionally square-root transformed.

  46. 46.

    Recorded after each UCS presentation by means of a seven-point Likert scale via microkeys.

  47. 47.

    Recorded with constant 0.5 V across two sintered-pellet 8 mm diameter AgCl electrodes filled with K–Y jelly from the hypothenar left-hand palm. FIR amplitudes (starting 1–4 s after CS onset) were range-corrected (Sect. 2.3.3.4 “Range Corrections”) using the largest UCR as the maximum range for each participant.

  48. 48.

    During the last 3 decades, EDRs have been frequently obtained in the context of startle eyeblink modulation (e.g., Böhmelt, Vanman, Dawson, & Boucsein, 1983; Lipp, Siddle, & Dall, 1997). Earlier research in this area focused on facilitation or inhibition of the startle response by a nonstartling stimulus presented prior to the startling stimulus and the influence of attentional factors on the magnitude of the eyeblink responses. Based on the bidirectional nature of the valence dimension of emotions, Lang, Bradley, and Cuthbert (1990) proposed that not only attentional but also emotional processes should have the ability of modifying the startle eyeblink. Since the nature of the startle response is more DR- than OR-like (Sect. 3.1.1.2), a positive emotional foreground stimulus should diminish the startle eyeblink magnitude, whereas emotional negative foreground stimulation should enhance its magnitude. Because of its indicator function for the intensity of especially negatively tuned emotional states, EDRs have come into use for verifying the nature of emotional foreground stimulation (e.g., Lipp et al., 1997). Since the startle eyeblink research does not primarily aim at EDR conditioning, it will not be reviewed in this section.

  49. 49.

    Recorded with standard methodology from the right hand’s hypothenar eminence. The amplitude criterion was set to 0.05 μS.

  50. 50.

    With two 1 cm diameter Ag/AgCl electrodes filled with K–Y gel (which is hypertonic) attached to the hypothenar palm of the left hand, the signal was visually inspected and corrected for artifacts. SCRs were scored if their maximum increase appeared within 1–4 s after CS onset. Zero responses were included in all analyses, and a range correction was applied using the largest UCR peak (between 9 and 13 s after UCR) in the acquisition phase as maximum individual value. A square-root transformation was applied to the SCR amp. data to improve distributional characteristics (Sect. 2.3.3.3).

  51. 51.

    This is a matter of evaluating superimposed EDRs (Fig. 2.16, Sect. 2.3.1.2 “Amplitudes of Exosomatic Responses Recorded with Direct Current”) but is not necessarily an argument in favor of discarding the distinction between FIR and SIR, as proposed by Pineles et al. (2009).

  52. 52.

    Recorded as SC with a Fels dermohmeter and zinc electrodes, taped to the first and third fingers of the participant’s right hand.

  53. 53.

    Recorded from the first and second left-hand fingers with constant current.

  54. 54.

    Recorded palmar/dorsal with zinc electrodes, saline electrode cream and a constant current of 40 μA. SRRs which occurred within 5 s after the presentation of the light were not scored. The NS.SRR freq. was transformed to reduce skewness and expressed as percentage of the NS.SRR freq. obtained during an initial 2-min resting period.

  55. 55.

    Measured as SR with a Fels dermohmeter, using 70 μA current, with zinc-manganese dioxide electrodes from the left palm and upper arm. Contact was made by a 5% saline-soaked cotton ball.

  56. 56.

    Recorded with Ag/AgCl electrodes and Beckman cream from ethanol-cleaned volar finger sites using 20 μA constant current. Red and green lights were used to signal that either an increase or a decrease of SRL would be reinforced by a flashing white light.

  57. 57.

    Roberts, Lacroix, and Wright (1974) could not observe instrumental conditioning of spontaneous SPRs in curarized rats. They used an electric shock as reinforcer which was applied when SPRs appeared exceeding either 10, 35, 60, or 75% of the greatest response during the baseline. As compared to a yoked-control group, no differences in SPR frequency appeared. Neither variation of shock intensity nor of curare dosage had an effect.

  58. 58.

    Recorded palmar vs. dorsal with 2 cm2 zinc electrodes filled with NaCl cream, transformed to log units.

  59. 59.

    Two SR channels with different sensitivity were used to avoid loss of SRR data in the high sensitivity channel (Sect. 2.1.3) which allowed the detection of 500 Ω changes (amplitude criterion). Eight-millimeter diameter zinc electrodes together with a zinc sulfate cream were placed on the participants’ fingers. Spontaneous EDRs were defined as not occuring within 6 s following any observable event which could give rise to an EDR (see Fig. 2.20, Sect. 2.3.2.2).

  60. 60.

    Recorded by Ag/AgCl electrodes and presumably hypertonic cream from the first and third fingers of the participant’s nondominant hand, voltage not reported.

  61. 61.

    Recorded with standard methodology except for using K–Y gel and a 1.0 V constant current. A Wheatstone bridge was incorporated to allow for greater amplification of the signal being transformed to SC. SCR amp. was calculated as the difference between the maximum SCL during 2–5 s and the minimum SCL during 0–2 s following the reward.

  62. 62.

    Recorded by means of a Wheatstone bridge, using 10 μA current, with a standard electrode/cream setting from the fingers. The amplitude criterion was 415 Ω for psychopaths and 376 Ω for controls.

  63. 63.

    Recorded with two-element lead electrodes, K–Y gel and 10 μA CC from the first and second fingers of the left hand, transformed to and stored as SCL, using a continuous tone pitch change as biofeedback signal.

  64. 64.

    SC reactivity was the peak change in SCL during the accident film, calculated as change from the last minute of the resting period.

  65. 65.

    These authors measured skin resistance with constant voltage (which normally gives conductance values) from the first and second fingers of the participant’s nondominant hand, using 20 × 25 mm electrodes.

  66. 66.

    Stern (1972) showed that SRR biofeedback training had only a small effect on the participants’ ability to detect whether or not they reacted electrodermally to a low-intensity sound. Furthermore, prior EDA biofeedback training was more effective for the detection of large as compared to small EDRs.

  67. 67.

    Recorded with dry silver electrodes taped to the palmar surface of the left index and middle fingers with unreported voltage; the biofeedback signal was presented visually on a sort of thermometer.

  68. 68.

    The reader is referred to reviews given by Raskin (1973), Spinks and Siddle (1983), or Koelega (1990), as well as Sect. 3.3.2.2.

  69. 69.

    Despite this component labeled “tonic” by Pribram and McGuinness (1975), the preparatory EDA is often clearly phasic in nature, as in the case of the SIR (Sect. 3.1.2.1). Additionally, the CNV paradigm, which is regarded by Pribram and collaborator as appropriate for testing preparatory activation, is mostly tested within an S1–S2 paradigm with intervals of less than 6 s. “Tonic” cannot be used here in the sense of longer-lasting shifts of arousal level; therefore, “phasic” would be more suitable. Since the sudomotor pathway used here is clearly connected to CNS structures preparing distinct motor actions (see Fig. 1.6, Sect. 1.3.4.1), the label “phasic” for this component is preferred here.

  70. 70.

    Recorded as SR unipolar between a 2 cm2 electrode at the middle finger and a 75 cm2 electrode on the upper arm with starch cream and 8 μA/cm2 constant current, transformed into SC scores.

  71. 71.

    Performed as a reanalysis, in terms of electrodermal recovery, of data reported by Furedy and Klajner (1972) from a study with high vs. low intensity and signaled vs. nonsignaled UCS.

  72. 72.

    This can be stated only for schizophrenic responders, since habituation cannot be measured in nonresponders (Sect. 3.4.2.2).

  73. 73.

    For the general excitatory and inhibitory role of the amygdala and the hippocampus in controlling EDA, see Sect. 1.3.4.1 “Subcortical Control of EDA".

  74. 74.

    The autonomic OR component is regarded as being elicited by “preattentive” mechanisms in Neisser’s (1967) sense.

  75. 75.

    Recorded as SR with standard methodology, using an additional abraded forearm site for grounding, transformed to SCRs, range-corrected with respect to the EDR following a 1 s, 100 dB white noise stimulus at the end of the experiment (Sect. 2.3.3.4 “Range Corrections”), and finally square-root transformed.

  76. 76.

    Recorded with standard methodology, detected by an analog computer using a slope criterion of 0.07 μS/s and an amplitude criterion of 0.3 μS.

  77. 77.

    There was no relationship between the occurrence of NS.SCRs and positive emotions, which points to the specificity of NS.EDRs as indicator for negative emotional states (Sect. 3.2.2.1).

  78. 78.

    Experiment 1 was performed with two independent groups (12 participants each of both genders), with vs. without omission in the 16th trial, either light-tone (1 kHz, 75 dB) or tone-light pairings as S1–S2. SCRs recorded with standard methodology were subjected to a range correction using a 100 dB stimulus at the end of the experiment to elicit a maximum SCR, and a square-root transformation.

  79. 79.

    Experiment 3 was performed with 48 participants, presenting light circles or tones as RT probes 1,300 ms following S2 omission and S2 re-presentation.

  80. 80.

    Similar stimulation and the same recording and evaluation techniques were applied as by Siddle and Packer (1987). A vibratory stimulus was used in addition to tone and light stimuli, all of them occurring equally often as S1, S2, and S3 in a Latin square design.

  81. 81.

    The various filter and channel capacity theories are summarized elsewhere (e.g., Broadbent, 1971; Massaro, 1975).

  82. 82.

    Recorded as SR (without reporting current density) with Beckman Ag/AgCl electrodes and K–Y gel from fingertips of the left hand, transformed to square-root conductance values. SCR amplitudes were evaluated quantitatively within 1–3 s following each critical word, contrarily to the Corteen group, which used an all-or-none amplitude criterion of 1 kΩ for an EDR appearing within a 13 s window.

  83. 83.

    Recorded from the left hand’s fingers with 1 cm2 Ag/AgCl electrodes and K–Y gel, within 1–4 s following stimulus onset. Two successive intervals without an SCR exceeding 0.02 μS were used as the habituation criterion (Sect. 3.1.1.3).

  84. 84.

    Including a square-root transformation, using a time window within 1.0–3.05 s following stimulus onset, and an amplitude criterion of 0.008 μS.

  85. 85.

    Clear trials were those without probes presented from 2 s before OR stimulus onset until 2 s after its offset.

  86. 86.

    Auditory tasks were the same as in the first experiment of Dawson, Filion, and Schell (1989). All tones presented to the to-be-ignored ear were designated task-irrelevant stimuli, whereas all tones presented to the to-be-attended ear were designated task relevant, regardless of their length.

  87. 87.

    Recorded with standard methodology from thenar/hypothenar left-hand sites; differences between mean SCLs during the task and the preceding baseline period were used as response measure which was logarithmically transformed (Sect. 2.3.3.3).

  88. 88.

    Recorded with a method described by Bechara, Damasio, Damasio, and Lee (1999), though incompletely reported, from thenar/hypothenar sites. SCRs appearing between the end of the reward/punishment window and the next decisive mouse click (on average 10 s) were considered anticipatory, whereas SCRs being generated after turning cards were considered reward or punishment SCRs. After removing tonic SCL changes by a difference-based transformation, an area measurement (see Sect. 2.3.1.4) was calculated and expressed in microseconds per time interval.

  89. 89.

    Recorded with constant 0.5 V from 1 cm2 Ag/AgCl electrodes attached to the medial phalanges of the nondominant hand’s middle and index fingers. Artifacts stemming from gross respiratory maneuvers (Sect. 2.2.5.2) were eliminated by means of the respiration signal obtained from a temperature sensor placed under the nostrils.

  90. 90.

    Recorded as stated in Footnote 88.

  91. 91.

    Alcohol-, cocaine/crack-, or metamphetamine-dependent inpatients shortly before the completion of their rehabilitation; 25 females, 21 males.

  92. 92.

    Abnormal EDRs in connection with poor IGT performance have also been observed in a variety of neurological and psychiatric conditions, e.g., substance and alcohol addiction (Bechara & Damasio, 2002; Bechara et al., 2001, 2002; Fishbein et al. 2005), Huntington`s disease (Campbell, Stout, & Finn, 2004), multiple sclerosis, Parkinson`s disease (Sect. 3.5.4.2), and anorexia nervosa (Tchanturia et al., 2007).

  93. 93.

    A possible LTM-modulatory mechanism for emotion-eliciting events, which includes influences from the amygdala and from hormonal systems on hippocampus-dependent declarative memory storage, backed up by PET studies and discussions of ANS responses in general, was suggested by Cahill and McGaugh (1998).

  94. 94.

    Recorded as SR by the use of a Wheatstone bridge, presumably using AC, transformed into log SCR. Ag/AgCl electrodes, saturated with 3% NaCl solution, were screwed into the arm-rest of the participant’s chair, fitting to the palm of the right hand.

  95. 95.

    Recorded with standard methodology, but using Beckman miniature electrodes and K–Y gel, within 1–5 s following stimulus onset, using an amplitude criterion of 0.024 μS, and square-root transformed.

  96. 96.

    Method of recording not reported. A range correction was performed.

  97. 97.

    It is, however, questionable whether geometric figures are really typical for right-hemisphere processing, since cognitive representation of abstract material may be more prone to left-hemisphere processing.

  98. 98.

    Using dry, bright-plated electrodes attached to the fingers by means of Velcro strap.

  99. 99.

    It is recommended using the same values as for the amplitude criterion (Sect. 2.3.1.2 “Choice of Amplitude Criteria”).

  100. 100.

    Amplitude criteria were 0.02 μS for SCRs and 0.2 mV for SPRs (monophasic from the SPL or biphasic from peak to peak). Latencies for SCRs were 1–5 s, and for SPRs 700 ms to 5 s after stimulus onset.

  101. 101.

    Recorded as SR using standard methodology (except a current density as high as 20 μA/cm2) from the distal phalanges of two fingers of each hand, transformed into SC values. Unfortunately, the EDA values reported in the results section remain ambiguous with respect to their unit and magnitude. The tasks were visual-imagery (15 slides, 9 of them with sexual content, with subsequent imaging); verbal-analytic (series of words, from which those denoting numbers had to be selected for calculation); auditory; and light stimulation.

  102. 102.

    Recorded between the distal phalanges of two fingers from both hands using standard methodology.

  103. 103.

    Recorded from medial phalanges of both hands as SR with standard methodology, transformed to SC and square-root transformed.

  104. 104.

    Recorded from 5 male participants with two Fels dermohmeters using a constant current of 70 μA each, with zinc/zinc sulfate electrodes and electrode cream (unspecified). Electrode positions were the center of each palm and the midline of the chest. Separate measures of SRL and mean NS.SRR amp. were taken during resting and during a serial learning task (as an appreciable stressor) on each of 24 or 36 days.

  105. 105.

    Recorded during three consecutive nights as NS.SPRs with standard methodology and Beckman cream, however with an unusually low time constant of 0.6 s (Sect. 2.1.4). Monophasic and diphasic SPRs (Sect. 2.2.3.1) exceeding 0.2 mV in at least one of the two hands were evaluated; differences less than 0.1 mV between both hands were not considered.

  106. 106.

    Recorded as SR (9.66 μA/cm2) with Ag/AgCl cup electrodes and 0.05 molar NaCl cream from the pretreated (rubbing with isopropyl alcohol and drying) volar middle phalanges of the first and middle fingers of both hands, transformed to SC.

  107. 107.

    Recorded with 1 cm diameter Ag/AgCl electrodes filled with 0.05% KCl solution in agar.

  108. 108.

    Older results in the area of EDA and arousal are reported by Duffy (1972) and Raskin (1973). More recent descriptions of activational, attentional, and cognitive phenomena with respect to their physiological concomitants can be found in the second volume of the series edited by Gale and Edwards (1983). A methodologically oriented, strongly generalized presentation of psychophysiological paradigms is given by Fahrenberg (1988).

  109. 109.

    In these experiments, EDA was recorded as SR from the soles of the feet with 2 × 4 cm lead electrodes and K–Y gel from acetone cleaned sites.

  110. 110.

    Recorded with dry lead electrodes from palmar finger sites against the forearm. A bipolar parieto-occipital EEG was recorded in parallel to the SR measurement.

  111. 111.

    However, Haider (1969) did not use the NS.EDR freq. but slow SP changes as an indicator of tonic EDA.

  112. 112.

    Furthermore, considerations of a differentiated view of the role of the RF in the elicitation of EDA were made by Sharpless and Jasper (1956), which could complement the neurophysiological concepts of EDA origins described in Sect. 1.3.4.1, if confirmed empirically. Those authors regarded the caudal (deeper) structures as the neurophysiological correlates of tonic EDA, while the rostral (higher) components of the RF were regarded as mainly contributing to phasic EDA phenomena, which indicate orienting or attentional processes (Sects. 3.1.1.1 and 3.1.3).

  113. 113.

    Fundamentally different neuronal trigger mechanisms of cardiovascular and electrodermal responses were also found in experiments with rats by Roberts and Young (summarized by Roberts, 1974). In a series of investigations into the effect of aversive stimuli upon approach behavior, consistent connections between HR and physical movement of rats were observed, which decreased over the trials, while both the SC and the negative SP component showed an ascending progress during the course of the trials. Roberts (1974) could exclude all possibilities of a somatic coupling of EDA, as in, for example, overall muscular tension or breathing, and he therefore presumed motivational and/or attentional processes (Sect. 3.1.3) as causative factors in the increase of EDA.

  114. 114.

    In the 1971 revision of his theory, Routtenberg became more careful with respect to the neuroanatomical structures that may underlie his “System II,” relating it closer to motor components of behavior.

  115. 115.

    To avoid making Fig. 3.4 more complicated, the thalamic parts of these loops are left out. Later, Alexander, Crutcher, and DeLong (1990) added another circuit which they identified as “limbic,” since it connects certain portions of the basal ganglia with structures of the limbic system, including the hippocampus, the amygdala, and the entorhinal cortex, in a very similar manner as motor and premotor or (pre)frontal cortical areas in the other two loops (or circuits, as labeled by Alexander et al., 1990). Connections of these limbic structures with the basal ganglia were left out in Fig. 3.4, since their primary function is supposed to be coordinating somatomotor concomitants of emotion-related mimics and gestures.

  116. 116.

    The close connections of the nigrostriatal dopaminergic fibers to motor behavior via the basal ganglia matches well with the slightly changed view of Routtenberg (1971) concerning his “System II” (see Footnote 115).

  117. 117.

    More recently, the basolateral amygdala has been identified as acting upon the the central amygdala to very rapidly eliciting fear responses via its projections to the hippocampus and the brain stem, whereas projections from the basolateral amygdala to the bed nucleus of the stria terminalis – together with those from the central amygdala – may be responsible for sustained fear responses, akin to anxiety (Heimer, Van Hoesen, Trimble, & Zahm, 2008). It is known from animal studies that the basolateral amygdala serves as input region, whereas the central amygdala plays a role in the mobilization of defensive behavior (Davis, 2000).

  118. 118.

    This is in accordance with Sokolov’s (1960) view of arousal within the OR (Sect. 3.1.1), insofar as an RF-mediated EEG desynchronization, together with hypothalamic ANS action patterns, being considered its concomitants.

  119. 119.

    From Latin cornu ammonis, which is another name for hippocampus. The hippocampus theta which is generated in the medial septal area, as already outlined by Routtenberg (1968), is transmitted to the CA3 field in the hippocampus to provide exact temporal information to the system. Gray (1982) suggested a feedback loop from the subiculum to the medial septal area, the anatomical confirmation of which is, however, lacking. Therefore it is not included in Fig. 3.4.

  120. 120.

    More recently, Gray and McNaughton (2000) made a distinction between fear, being mediated by the flight/fight freeze system, and anxiety, being mediated by the BAS.

  121. 121.

    Gray (1982) explained the anxiolytic properties of tranquilizers as well as hypnotics via their facilitating properties on the inhibitory action of GABA (gamma-aminobutyric acid) on noradrenergic and serotonergic synaptic transmission. Tranquilizers of the benzodiazepine type act directly via a specific postsynaptic receptor, while hypnotics like barbiturates and alcohol act indirectly through a blockade of picrotoxine receptors, a substance which inhibits GABA.

  122. 122.

    Gray (1987, p. 226) pointed to the difference between his own and Panksepp’s (1982) view concerning fear and anxiety (Sect. 3.4.1.1), the latter identifying those largely with the activity of the fight/flight system, while Gray identified them with BAS activity.

  123. 123.

    Those neurons, which are not depicted in Fig. 3.4, receive afferent impulses from the medial hypothalamic decision center (Gray, 1987, p. 265).

  124. 124.

    Fowles (1988) reported data from a doctoral dissertation performed by Fisher in 1985, examining the effects of 10, 50, and 100% success using 20 participants per group. In contrast to a feedback-only condition, a monetary incentive condition, holding the amount of money earned constant across the different success groups, yielded significantly greater HRs regardless of the amount of success. On the other hand, NS.SCR freq. was significantly heightened in the 10% group as compared to the other ones, regardless of monetary incentive, which supports EDA not being influenced by appetitive motivational states during task performance. Sosnowski, Nurzynska, and Polec (1991) could not find an influence of manipulating monetary reinforcement on both HR and SCR amp. (recorded with standard methodology) in 60 student women randomly assigned to reward, frustration, and control groups performing a problem-solving task. Participants were run in pairs; the active participant solved the problem, being observed by the passive one. During the task, SCR amplitudes decreased significantly in active participants while being markedly increased in passive participants. The authors interpreted this result as being in accordance with Fowles’ EDA-BIS hypothesis since EDA was increased in what they called passive coping.

  125. 125.

    In her study comparing HR-reactive and HR-nonreactive participants under various stimulation conditions, Lawler (1980) found that HR reactivity being associated with minimal EDA, while HR nonreactive individuals showed more NS.SCRs.

  126. 126.

    Altogether, the BIS-EDA connection seems to be better established than the BAS-HR connection (see as an example, Gomez and McLaren’s (1997) study described in Sect. 3.1.2.2).

  127. 127.

    However, the postulated relationships were not always confirmed. Brenner, Beauchaine and Sylvers (2005) performed a study with 50 participants (28 females, 22 males) who were given different reward and nonreward conditions in sequence during a number-recognition task, while cardiovascular and electrodermal measures were recorded as markers for BIS/BAS activity. Unexpectedly, electrodermal reactivity was elevated not only during nonreward but also during reward. Furthermore, the proposed psychophysiological markers for BIS and BAS did not yield substantial correlations with a BIS/BAS questionnaire but only with subjective measures of affect.

  128. 128.

    This result is at variance with the Neumann (1968) findings described in Sect. 2.4.1.1 who obtained different SRL patterns in summer and in winter. There were also seasonal differences in EDA in the Mauritius study (Sect. 2.4.1.1), since Venables and Mitchell (1996) reported a significant interaction between gender and season, yielding a greater responsivity of females in hot weather.

  129. 129.

    However, in their review on the indicator function of EDA for arousal and emotion, Sequeira, Hot, Silvert, and Delplanque (2009) came to the conclusion that EDA might be used as a valid indicator for diurnal variations in emotional reactivity, which is not necessarily represented in subjective emotionality.

  130. 130.

    Unfortunately, for technical reasons such as electrode detachment, only 12 of the 24 participants (7 with the horror tape, 5 controls) could be included in the sleep evaluation. As could have been expected, tonic EDA increased during listening to the horror story compared to the boring story. No differences were found in HR.

  131. 131.

    Recorded with Ag/AgCl sponge electrodes from the fingers of the right hand with an “inert” electrolyte and 21 μA/cm2 for SR, and from the left middle finger to the scrubbed left forearm with Redux cream, and 1 s time constant, for SP. Amplitude criteria: 50 Ω and 100 μV (positive or negative), respectively.

  132. 132.

    Recorded with silver cup electrodes and an EEG electrode cream from acetone-cleaned sites, each pair having interelectrode distances of 5–6 cm. They were placed palmar/dorsal, at the dorsal forearm; and over the frontal, trapezius, deltoid, biceps, extensor digitorum, or other muscles, approximately along the neuraxis and peripheral nerves.

  133. 133.

    Recorded with Beckman electrodes, time constant 0.6 s, low-pass filtered with 15 Hz, amplitude criterion 200 μV. Results were log-transformed.

  134. 134.

    Pivik (1978) reported that individuals being asked to get more involved as usual in their own dreams showed an increase of EDA during sleep.

  135. 135.

    Recorded from comparable unipolar right and left sites (thenar against forearm) with Ag/AgCl sponge electrodes; amplitude criteria 200 μV and 100 Ω, respectively.

  136. 136.

    SRL and SRRs measured from different washed volar sites at two fingers, with Ag electrodes, 5 mm in diameter contact area. An AC-coupling circuit was used that resulted in biphasic SRRs.

  137. 137.

    DC-recorded thenar and dorsal against a skin-drilled forearm site with Ag/AgCl electrodes and 0.05 M NaCl agar-agar cream. Amplitude criterion for SPRs (negative, diphasic, and positive waves) was 250 μV.

  138. 138.

    Recorded from the left middle finger by Ag/AgCl electrodes, NaCl cream, and 10 μA/cm2 constant current. Amplitude criterion for NS.SRRs: 50 Ω.

  139. 139.

    Recorded with Beckman Ag/AgCl electrodes as SR from the fingers of the right hand using 40 μA, and as SP between left index finger and forearm with 0.24 s time constant. A time window of 13 s following stimulus onset was used to determine EDR; SPR amp. measured as the total biphasic amplitude (Sect. 2.3.1.2 “Amplitudes of Endosomatic Responses”); SRR amp. being transformed to log SCR amp.

  140. 140.

     kHz, 30 dB tones presented over a loudspeaker for 45 s. Amplitude criteria: 125 Ω for SRRs and 100 mV for SPRs, within a time window of 1–5 s after stimulus onset.

  141. 141.

    Recorded between scrubbed volar forearm and thenar sites with a 0.3 s time constant. A 1 kHz, 1 s tone was presented over loudspeakers with 70 dB at the participant’s head, using three regular (10, 20, and 30 s) and three randomized irregular ISI conditions (8, 10, or 12; 16, 20, or 24; and 24, 30, or 36 s).

  142. 142.

     Hz, 75 dB, 1 or 2 s duration, at 30, 45, and 60 s ISIs. SRRs were obtained by dividing the prestimulus SRL by the maximum poststimulus resistance change occurring within 7 s after stimulus onset, log-transformed.

  143. 143.

    In a famous series of field studies, differences in time courses between electrodermal and subjective measures of stress had been obtained within nonlaboratory research with parachutists (for a summary, see Epstein, 1972). Whether or not psychophysiological dissociations during the course of anticipating stressful events will increase with the complexity of the stress situation should be further investigated. This requires combined laboratory/field studies in the area of stress research, which are also necessary to establish possible relationships between short-lasting psychophysiological stress effects and long-lasting changes.

  144. 144.

    With respect to the theories of the psychobiology of emotions in general, the reader is referred to appropriate comprehensive reviews (e.g., Grings & Dawson, 1978; Panksepp, 1986; Plutchik, 1980; Schwartz, 1986).

  145. 145.

    Recorded with 600 Hz, 1 V, using an RC bridge (Sect. 2.1.5) and metal electrodes with 1 cm diameter. SZR amp. were subjected to log transformation with respect to a suggested validity of Fechner’s law for EDR.

  146. 146.

    Recorded by the use of a 60 Hz AC and a bridge from volar finger surfaces, evaluated as SY, using 1 μS as amplitude criterion.

  147. 147.

    Uchiyama (1992) brought real-life situations in the laboratory for inducing the emotions fear, anger, and joy in 6 male participants. NS.SCR freq. (recorded with unreported methodology) was found to be higher in the fear situation (i.e., informing the participant that he would suffer from a heart disease) compared to the others.

  148. 148.

    Ax (1953) did not obtain measures of catecholamines but concluded that the ANS pattern under anger resembled the expected response to adrenaline and noradrenaline injections, while the ANS pattern under fear resembled the response in case adrenaline would have been injected.

  149. 149.

    Fear was induced by tape presentation of a fear-evoking short story being accompanied by an unannounced darkening of the room; anger was induced by presenting a series of anagrams developed by Boucsein and Frye (1974), most of which were unsolvable; however, the insoluble nature of the anagrams was not detected by the participants during their presentation; and pleasure was induced by positive reinforcement and the announcement of increased payment at the end of the study.

  150. 150.

    Recorded with standard methodology from the left hand’s fingers. To obtain an additional objective measure of the “forehead anxiety sweat,” another EDA recording was taken from the forehead. Electrodes were fixed by means of histoacryl (Sect. 2.2.2.1); the evaluation of EDA followed Thom’s (1988) method (Sect. 2.2.4.2).

  151. 151.

    A further domain is animal research about which Panksepp (1982, p. 410) stated that brain research in this area seems to regard the study of emotional expression as the only credible scientific approach.

  152. 152.

    As inferred from another article of the Lanzetta group (Kleck et al., 1976), EDA was probably recorded with 3.14 cm2 zinc electrodes using a zinc sulfate electrode cream and a Fels dermohmeter with 70 μA constant current. SRL values recorded from palmar sites were transformed to SCL values. The anticipatory EDR was computed as an increase from the average SCL 2 min before and at the beginning of shock application to the average SCL from 2 min before and at the beginning of the slide projection. The latter SCL was used as a reference for the EDR to shock application, thus being subtracted from the average SCL 6 and 8 min after its application.

  153. 153.

    The relationships between voluntarily expressed facial emotions and EDA (among other ANS variables) have been studied in 20 elderly individuals (71–83 years) of both genders by Levenson, Carstensen, Friesen, and Ekman (1991). The magnitude of ANS changes was smaller compared to young individuals.

  154. 154.

    EDA was recorded as SR via palmar Ag/AgCl electrodes, transformed into SC and individually standardized.

  155. 155.

    Gross and Levenson (1993) investigated the effect of suppressing emotional expression while watching a short disgusting film, using the SCL, various cardiovascular and respiratory measures, somatic activity, and eye blink frequency as psychophysiological indicators. Suppression reduced the expressive behavior in their 42 female and 43 male participants but did not reveal a clear-cut psychophysiological pattern. Interestingly, participants in the suppression condition showed significantly higher increases in SC than their nonsuppression counterparts, but also yielded significantly greater HR deceleration. This was interpreted by the authors as reflecting the possibility that voluntary suppression of facial disgust expressions may lead to greater physiological “disgust-like” responding (Gross & Levenson, 1993, p. 981). Gross (1998) proposed an antecedent- and response-focused model of emotion regulation based on the work of the Lazarus group reported in this section, which he supposed having beneficial consequences for health. For testing his model, Gross presented 120 participants (60 of each gender) stressful films. Among other measures, SCL was recorded during baseline, instructions, film presentation, and a postfilm period. After the first presentation of either a stress inducing or a neutral film, a second film was announced with the instruction to simply watch, reappraise, or suppress during the presentation (40 participants per group). As in the Gross and Levenson (1993) study, suppression increased SCL markedly, compared to both baseline recordings and the other two groups. The author concluded that the health-promoting mechanism suggested here might result from an increase of the classical sympathetic stress response which may in turn influence the course of immune responding.

  156. 156.

    Unipolar recordings with zinc electrodes and a zinc sulfate electrode cream from palmar against forearm sites, using a “low” constant current, thereafter transformed into SC values.

  157. 157.

    This social learning should have appeared early during life, since Buck (1977) found comparable negative correlations between SCR and communication accuracy even in preschoolers, giving nonverbal messages via spontaneous facial expressions and gestures to their mothers.

  158. 158.

    Gender differences in expressed emotions were observed by Kring and Gordon (1998) who presented 22 female and 21 male participants six brief emotion-inducing film clips (two of each: happy, sad, and fear), separated by neutral film segments. SC was recorded with standard methodology from the thenar eminence of the nondominant hand, and videotaped facial expressions were automatically coded. NS.SCR freq. was highest for fear in both genders, whereas females yielded higher values in the sad category than males, the reverse of which was found for happy film contents. The EDA results did not correspond to the evaluation of facial expressions.

  159. 159.

    With standard methodology, using K–Y gel. Mean SC change was calculated by subtracting the 1 s prestimulus average from the average between 2 and 7 s after stimulus onset.

  160. 160.

    Recorded with standard methodology from distilled water-cleaned medial phalanges of the third and fourth fingers of the right hand. SCR amp. were obtained 1–4 s after stimulus onset with an amplitude criterion of 0.05 μS.

  161. 161.

    It is not obvious that surprise does predominantly elicit positive emotions. If strong enough, it may as well elicit a DR which is normally negatively tuned (Sect. 3.1.1.2).

  162. 162.

    Combining EDA with fMRI recordings, which has been recently available (Sect. 2.2.3.5), together with mathematical solutions for the evaluation of overlapping EDRs, constitutes an intriguing new possibility of research into CNS correlates of processing emotion-expressing faces. This had been demonstrated by Williams et al. (2004) in their study described in Sect. 2.3.1.5. Their fMRI data revealed that the U-shaped SCR pattern across early, middle, and late presentations of face stimuli expressing fear was paralleled by a temporal sequence of somatosensory insula, dorsomedial prefrontal cortical, and left amygdalar activation.

  163. 163.

    Ekman et al. (1985) in an experiment on startle simulation and suppression revealed that startle should be considered a reflex rather than an emotion. To the present author’s view, the widely used paradigm of startle response modification by manipulations of the emotional background might be regarded as a sideline in emotion research which concentrates on a single dependent measure. Revealing its contribution to the core of psychophysiological research into emotional states may require further theoretical integration.

  164. 164.

    An attempt to determine an individual’s most reactive ANS channel (i.e., EDA or HR) was made by Levis and Smith (1987), using the balloon-burst test to preclassify their participants as high SC responders, high HR responders, high responders in both channels, or low responders in both channels. In a subsequent presentation of a fear-eliciting slide (a man who died in an accident), those participants found to be high responders on a given channel showed greater reactivity on that particular channel as compared to low responders.

  165. 165.

    Seligman (1975), in an explorative study with 6 participants, obtained differential effects of pleasant and released vs. unpleasant and inhibited feelings (as reported on a Mood Adjective Check List) on negative vs. positive SPR waves, respectively, during ten counseling sessions of 50 min duration each. However, Edelberg (1972a) already pointed to results with respect to the emotional valence of different SP wave forms being equivocal in general.

  166. 166.

    Recorded with a Fels dermohmeter and 70 μA current, by means of zinc/zinc sulfate electrodes (cream not mentioned), transformed into SC units.

  167. 167.

    Bilateral thenar recording as SR with Beckman electrodes of 1 cm diameter, 10 μA constant current, transformed to SC values.

  168. 168.

    Recorded with 10 μA constant current from thenar/hypothenar sites, using Ag/AgCl electrodes of 1 cm diameter with K–Y gel, amplitude criterion 80 Ω. Responses in 10-s intervals were averaged and transformed to log SC.

  169. 169.

    Recorded with Ag/AgCl sponge electrodes (other details of recording not reported) as SRL values, averaged for every 10 s and transformed to SCL values.

  170. 170.

    Crude surgeries on the male genitals of a primitive native culture in Australia. EDA has been recorded from both hands as SR with zinc electrodes of 2 cm diameter, using an agar–zinc electrode cream and a 40 μA constant current. SR values were transformed to log SC changes (differences between pre- and postaccident SCLs).

  171. 171.

    A phylogenetically oriented approach into the psychophysiological responses to film stimuli has been performed by Westbury and Neumann (2008) with 73 participants (37 females, 36 males). Four 10-min clips for each of five target groups (depicting humans, primates, quadruped mammals, and birds in victimized circumstances) were presented in randomized order. Participants were asked to rate their empathy, SC (recorded according to some of the standards, but with “surgicon” electrolyte paste), respiration and corrugator muscle activity were recorded. SCRs were obtained as trough-to-peak distance within 1–4 s after the film presentations and square-root transformed (Sect. 2.3.3.3). Both SCRs and empathy ratings increased with phylogenetic similarity to humans across animal groups, being highest in the films showing human victims.

  172. 172.

    Recorded as SR with dry 2 cm2 lead electrodes from previously acetone-cleaned finger sites, transformed to values of μS/cm2.

  173. 173.

    Measured unipolar from left middle finger against forearm sites with electrodes (metal not specified) of 0.32 cm2 and 3 × 4 in., respectively, starch cream and 20 μA constant current. 100 Ω were used as amplitude criterion.

  174. 174.

    Recorded unipolar thenar vs. forearm with 5 and 58 cm2 Ag/AgCl electrodes, respectively, filled with 0.5 M NaCl (probably liquid) electrolyte, using a current of 10 μA/cm2. Amplitude criterion for NS.SRRs was 100 Ω.

  175. 175.

    Measured with Ag/AgCl electrodes of 4 mm diameter, KCl cream (0.67 M) in agar, and 0.5 V, evaluated as log SC change from pretreatment level.

  176. 176.

    Recorded as SR with 10 μA constant current, Beckman Ag/AgCl electrodes, and K–Y gel palmar/dorsal. Separate channels were used for SRL and SRRs (sensitivity 50 Ω). SRL values were transformed to SCL, and SCRs were computed as differences between logarithmized SCLs before and the maximum within 6 s after stimulus onset.

  177. 177.

    Recorded as SR between palm and forearm with Beckman electrodes and cream, transformed to square-root SCR. Time window for SCRs: 0.5–3 s after stimulus onset; amplitude criterion for NS.SRR freq.: 200 Ω.

  178. 178.

    This could be shown by Overmier (1985) with animals, using completely different experimental conditions and plasma cortisol as stress indicator.

  179. 179.

    Recorded palmar with Ag/AgCl electrodes and 10 μA constant current, transformed to SC. No information was provided concerning electrode size, cream, and amplitude criterion for SCRs.

  180. 180.

    The term “strain” comes from physics, where strain is used for the response of a system being under stress.

  181. 181.

    Recorded with Ag/AgCl electrodes of 6 mm diameter from previously washed sites on the medial phalanges of the index finger and thumb of the nondominant hand with hypoallergenic gel and a CC below 0.5 V.

  182. 182.

    See also the Dindo and Fowles (2008) study which is described in Sect. 3.1.1.1.

  183. 183.

    Eysenck discussed his inhibition concept with respect to Pavlov’s “transmarginal inhibition” or “protective inhibition,” the neurophysiology of which he regarded as unrealistic, though the phenomenon itself (which is also in accordance with the so-called Yerkes–Dodson law, and the inverted U-shaped relationship between arousal and performance; see Sect. 3.2.1.1) has been frequently observed (Eysenck, 1983, p. 18). However, Eysenck’s attempt to include different aspects of inhibition into a unitary CNS inhibition concept may have contributed to various differences between theoretical concepts and experimental results (Nebylitsyn, 1972, p. 21; Strelau, 1983, p. 145), thus being in part responsible for pitfalls in establishing reliable psychophysiological correlates of extraversion/introversion.

  184. 184.

    Recorded with zinc electrodes of 25 mm diameter and 1% zinc sulfate cream, current density not reported.

  185. 185.

    Recorded with 2 cm2 Ag/AgCl electrodes, 0.5% KCl in Unibase, and 1.0 V constant current.

  186. 186.

    Recorded with Ag/AgCl electrodes of 1 cm diameter, 0.05M NaCl cream, and 9.55 μA/cm2 current density; SR values transformed into SC values. SCLs and SCRs were obtained at the beginning and after each test stimulus. SCR amp. were square-root transformed and range-corrected.

  187. 187.

    Indirect support comes from an investigation of Wilson (1990) with 61 male and 50 female participants who self-recorded their SCL throughout a whole working day. The recording device featured dry electrodes being not fixed to the skin which constituted a somehow instable methodology. By applying various questionnaire measures, Wilson came to the conclusion that sociability, rather than impulsivity, seemed to have been the component of extraversion responsible for individual variations in the diurnal SCL curve.

  188. 188.

    Recorded with standard Beckman Ag/AgCl electrodes and Beckman cream from the left palm, using 20 μA/cm2 as current density and 100 Ω as an amplitude criterion.

  189. 189.

    Such a contrary to classic expectation of greater electrodermal reactivity of low trait anxious compared to high trait anxious individuals has, under conditions of relatively low stress, been more recently confirmed by Wilken, Smith, Tola, and Mann (2000) and by Naveteur, Buisine, and Gruzelier (2005). Wilken et al. tried to explain their observation with the inverted-U-shaped relationship between general arousal and behavioral outcomes (Sect. 3.2.1.1). Trait anxious participants might have come to their experiment with higher arousal than their low anxious counterparts, thus reaching the peak of the inverted-U arousal curve earlier. It remains, however, unclear why EDA as a direct measure of general arousal should decrease with a further increase of arousal after this peak. Their alternative interpretation with respect to the LIV (Sect. 2.5.4) seems to be more appropriate here.

  190. 190.

    Individual ANS reactivity was obtained by using the higher of either HR or EDA standardized scores. However, since EDA was used in all six studies but not HR, individual reactivity might have been mainly expressed within the electrodermal system. EDA was recorded in all studies as SR by means of a Fels dermohmeter with a constant current of 70 μA across palmar zinc/zinc sulfate electrodes with 2 cm diameter.

  191. 191.

    Recorded with 16 mm Ag/AgCl electrodes and 0.05 M NaCl in Unibase from the palms. Amplitude criterion: 100 Ω. Reaction scores were corrected for artifact-free baselines.

  192. 192.

    Recorded as SR with Ag/AgCl electrodes from the medial phalanges of the index and middle fingers of the nondominant hand with 8 μA CC, transformed into SC, amplitude criterion 0.02 μS.

  193. 193.

    The authors performed unusual SR recordings with an active palmar and an inactive forearm electrode, Beckman standard electrodes and cream, and 20 μA constant current. SR scores were transformed to SC, and SCR amplitudes were square-root transformed.

  194. 194.

    Recorded as SR with Ag/AgCl electrodes and NaCl electrode cream from Beckman, unipolar palmar against forearm, using 20 μA/cm2 current, performing transformation to SC and logarithmization. The amplitude criterion for NS.EDRs during the resting condition was 100 Ω.

  195. 195.

    Measured with nonpolarizing Ag/AgCl sponge electrodes between the palm and an alcohol-cleaned forearm site, AC-coupling (Sect. 2.1.3) with 0.45 s time constant, and an amplitude criterion of 0.1 mV. Several fluctuations within a 6 s window were regarded as a single SPR.

  196. 196.

    For reliabilities of frequencies and mean amplitudes of NS.EDRs, see sections 2.5.2.1 under “Tonic Skin Conductance Measures” and 2.5.2.2 under “Tonic Skin Resistance Measures”.

  197. 197.

    NS.SCRs were recorded with standard methodology using a Wheatstone bridge (Sect. 2.1.3) during 40 s after a 5 min relaxation period, using an amplitude criterion of 0.05 μS (Sect. 2.3.1.2 “Choice of Amplitude Criteria”).

  198. 198.

    SR recorded palmar against a forearm site with 2 cm2 Ag/AgCl electrodes, Beckman cream, and 10 μA/cm2 constant current. Amplitude criterion: 100 Ω.

  199. 199.

    SCRs were recorded with standard methodology and an amplitude criterion of 0.02 μS during a 5 min period of no stimulation.

  200. 200.

    The possibility of electrodermal lability in adolescence being a protective factor against criminal behavior during adulthood since it prevents underarousal has been discussed by Raine, Venables, and Williams (1995); see also Sect. 3.4.1.2. As a possible explanation, Crider (2008) suggested an increased amount of self-control of emotional behavior in electrodermal labiles, which would be in line with his effortful control hypothesis.

  201. 201.

    Crider (1993) stated that the temporal stability of electrodermal lability measures is somewhat lower than usually demanded of measures for individual differences.

  202. 202.

    Typical applications of EDA within therapeutic techniques used in psychopathology can also be found in biofeedback (Sect. 3.1.2.3) and in systematic densitization (summarized by Katkin & Deitz, 1973).

  203. 203.

    Lader and Wing’s research had been continued by Chattopadhyay and co-workers Chattopadhyay, 1981; Chattopadhyay et al., 1975, 1980, 1982, 1983.

  204. 204.

    Recorded with 10 μA CC using lead electrodes, filled with 0.05 M NaCl cream, from the distal phalanx of the right thumb against an inactive (rubbed) site on the lateral aspect of the arm above the elbow. SR was converted to log SC, and an SCR of 0.003 log μS was used as amplitude criterion (Sect. 2.3.1.2 “Choice of Amplitude Criteria”).

  205. 205.

    Possibly the small difference in amplitude criteria, 0.002 log μS used by Hart (1974) vs. 0.003 log μS used by Lader and Wing (1964), may also have contributed to that difference (Sect. 2.3.1.2 “Choice of Amplitude Criteria”).

  206. 206.

    With standard methodology from two electrodes adjacently placed to the participant’s right-hand hypothenar eminence, with a resolution of 0.001 μS, a sampling rate of 10 Hz, scored as SCL difference between the mean during the 3 s picture viewing and a baseline, log-transformed and range-corrected (Sects. 2.3.3.4 “Range Corrections” and “Transformation into Standard Values”).

  207. 207.

    Recorded from the middle phalanges of digits 3 and 4 of the left hand with disposable Ag/AgCl electrodes of 2 cm2 contact area and an isotonic electrode cream, using 0.5 V constant voltage and a 6 Hz sampling rate. The SCL was controlled for drift due to progressive skin moistening and/or maceration (Sect. 2.2.6.1) by simultaneous application of new electrodes after the flight. No adjustment was made since the SCL downward drift did not exceed 8.2% on average. Artifacts were removed during the interactive signal evaluation (Sect. 2.3.4.1).

  208. 208.

    Recorded with standard methodology from the middle phalanges of digits 3 and 4 with 2 cm2 disposable Ag/AgCl electrodes. Amplitude criterion 0.02 μS.

  209. 209.

    SRL, NS.SCR amp., and mean habituation rate discriminated between patients having anxiety disorders with and without panic attacks (Birket-Smith, Hasle, & Jensen, 1993). During anticipation of delivering a speech and other stressors, patients with panic disorders showed substantially higher SCLs and NS.SCRs compared to healthy controls (Braune, Albus, Fröhler, Höhn, & Scheibe, 1994). Compared to healthy controls, panic patients yielded significantly higher NS.SCR freq. during a series of performance tests (Dratcu & Bond, 1998).

  210. 210.

    Another anxiety-related disorder in which EDA has been used for probing conditioning is posttraumatic stress disease (e.g., Orr et al., 2006).

  211. 211.

    Since SCL data were not collected for 16 participants during their diagnosis and subsequent treatment, the results are based only on 31 participants. The recording method was not explicitly mentioned.

  212. 212.

    Recorded with CV of about 0.5 V, using gold-plated dry electrodes attached by plastic tape to the first and third finger of the right hand. The volar finger sites were soap-washed, dried, and alcohol cleaned prior to attachment.

  213. 213.

    SCR amp. was obtained as difference between the SCL at the peak and the SCL immediately before SCR onset.

  214. 214.

    Recorded with 7 mm diameter zinc electrodes and zinc sulfate electrolyte, using a Wheatstone bridge. The polarity of the 1.5 V reference was reversed every 1.2 s.

  215. 215.

    Bilateral recordings with Ag/AgCl electrodes of 4.5 mm diameter, filled with 0.5% KCl in agar-agar cream, using 0.5 V constant voltage. Amplitude criterion: 0.05 μS.

  216. 216.

    Although schizophrenia has been the domain for electrodermal nonresponding, this phenomenon is clearly not specific to schizophrenia (Iacono, 1985; see Sect. 3.4.2.2). Raine, Bihrle, Venables, Mednick, and Pollock (1999) assessed SC-OR and schizotypy in 134 males during adolescence. Criminal offending and alcohol abuse were assessed 12 years later. Schizotypic individuals who became criminals had been characterized by deficits in ED-OR during adolescence.

  217. 217.

    The term “recovery rate” used in this context does not correspond to the same term used for recovery speed, which is the reciprocal of EDR recovery time (see sections 2.3.1.3 under “Recovery Parameters” and 2.3.3.3).

  218. 218.

    EDA was recorded with Ag/AgCl electrodes of 9 mm diameter with 0.9% NaCl cream from sites of the left hand’s second and third finger, prior cleaned with an ethanol (75%)/acetone (25%) mixture. The authors used a combined current density (9 μA/cm2)/voltage (2.7 V) limiting system. If necessary, the SCR rec.t/2 values were extrapolated by the method of curve matching (Sect. 2.3.1.3 “Recovery Parameters”). Since the mean recovery time of the EDR to the first tone was conspicuously longer than those following the other tones, the latter ones were averaged, and analyses were performed separately for the first recovery and the mean of the others. The amplitude criterion for NS.SCRs was 0.0043 log μS.

  219. 219.

    Lower electrodermal activity/responsivity was also demonstrated in disruptive children (van Goozen, Matthys, Cohen-Kettenis, Buitelaar, & van Engeland, 2000) and for boys with conduct disorder (Herpertz et al., 2005).

  220. 220.

    Recorded bilaterally from the medial phalanges of the second and third fingers with 0.45 cm diameter Ag/AgCl electrodes and 0.5% potassium chloride in 2% agar-agar as electrolyte. Amplitude criterion 0.05 μS.

  221. 221.

    The independence of this sample from the one used by Raine et al. (1990a) remains unclear.

  222. 222.

    Such a motivational imbalance was also mentioned as a key factor in psychopathy by Arnett (1997) in his review on ANS responsivity in psychopaths.

  223. 223.

    Fung et al. (2005) observed reduced anticipatory SCRs in 65 psychopathy-prone 16-year-old males compared to 65 controls, demonstrating that the anticipatory electrodermal hyporesponsivity of psychopathic adults is an ANS impairment which may be already present in adolescents and could predispose individuals to adult psychopathy.

  224. 224.

    Damasio, Tranel, and Damasio (1990) observed reduced EDA to social stimuli in 5 patients with sociopathic behavior who had lesions in the VMPFC (Sect. 3.1.3.3), in comparison to six brain-damaged and five normal controls who had no diagnosis of sociopathy. In a study combining fMRI with EDA, Raine, Lencz, Bihrle, LaCasse, and Colletti (2000) found both damage to gray and white matter in the prefrontal cortex plus lowered SC under stress compared to healthy controls in 21 male neurological patients who developed a pseudopsychopathic personality, thus relating subtle structural deficits in the prefrontal cortex to antisocial behavior and electrodermal hyporeactivity.

  225. 225.

    Recorded as SR with a constant current of 6.4 μA/cm2, using Beckman Ag/AgCl electrodes filled with K–Y gel. Recordings were performed during 5 min of resting, a free word-association test, stimulation with tones and bells of moderate intensity, a differential classical conditioning, and an RT task. EDA was expressed in terms of SC parameters.

  226. 226.

    As an exception, EDA has also been found to be higher in depressive patients than in controls. In a study with 18 patients suffering from major depression (9 of each gender), who had not taken any medication for at least 10 days, and 18 controls (6 females, 12 males), Guinjoan, Bernabó, and Cardinali (1995) reported significantly higher NS.SPR amp. in patients while raising from lying to standing and during handgrip exercise, but not during other exercises such as hyperventilation, cold pressor test, and mental arithmetic. The authors interpreted their results being due to an altered sympathetic/parasympathetic balance in depressives.

  227. 227.

    Recorded with 9.5 mm diameter double-element lead electrodes unipolarly (abraded arm site above the elbow vs. thumb), using a 0.05 M NaCl electrolyte and 14 μA/cm2 constant current.

  228. 228.

     kHz, 1 s tones of either 85 or 105 dB; 10 and 12 stimuli, respectively; ISI 20–40 s. SC was recorded bilaterally with standard methodology; SCL measured immediately before each tone and averaged; SCR was obtained within 1–3 s after stimulus onset, with an amplitude criterion of 0.05 μS.

  229. 229.

    Recorded with 0.5 V CC and transformed to in μS/cm2.

  230. 230.

    The type of EDA was not specified; 8 mm diameter Ag/AgCl electrodes filled with conductive cream were attached to the medial phalanges of both index and middle fingers.

  231. 231.

    Recorded with 7 mm diameter zinc electrodes, sponges saturated in zinc sulfate solution used as electrolyte, sites washed and cleaned with alcohol.

  232. 232.

    Venables (1983) mentioned several critical points in the investigations of the Mednick group. First, the ISIs in their classical conditioning paradigm may have been too short to allow separation of the different kinds of EDRs with respect to their latencies (Sect. 3.1.2.1). Second, the differences found in SCR lat. may have been confounded with different absolute auditory thresholds for different frequency ranges, as can be observed in schizophrenics.

  233. 233.

    Mednick (1978) pointed to different selection procedures: the New York study excluded individuals from nonintact families, while the Copenhagen study did not.

  234. 234.

    Patterson (1976) detected 11 nonresponders (Sect. 3.4.2.2) among 31 male chronic schizophrenics. The remaining 20 participants showed a bimodal distribution with respect to SCR rec.t/2, and the fast recovery participants showed significantly slower pupillary constriction in the light/dark reflex as compared to the slow recovery participants, which was discussed by the author as possibly due to a greater adrenergic outflow in the first group. In addition, these results also question the generalizability of the results that shorter EDR recoveries appear in schizophrenics.

  235. 235.

    As can be inferred from other publications of the Zahn group, zinc/zinc sulfate electrodes with 0.79 cm2 area were attached to palmar sites.

  236. 236.

    It could be further speculated that different EDA parameters are related to different pathogenetic factors. Cannon et al. (1988), using a subsample from the Copenhagen study, found that individuals with enlarged third ventricles (which may point to hypothalamic and/or amygdalar deficits) showed significant overall reductions in EDR amp. and in percentage of EDRs observed during OR and conditioning trials (see also Sect. 3.4.2.2).

  237. 237.

    About 3 years after the Copenhagen study, an Israeli high-risk study started with 100 children at an average age of 11 years, half of them being offspring from schizophrenic parents, the other half not. All were subjected to an SC-OR paradigm. At about 26 years of age, 90 participants were reevaluated, 27 of whom (mostly from the high-risk group) were diagnosed as schizophrenic according to DSM-III. Results of the differences in SC-ORs between the groups and additional results from an investigation of the children at the age of 16 were reported by Kugelmass et al. (1995). To-become schizophrenics had significantly higher anxiety ratings at age 16 than the nondiagnosed participants. Unexpectedly, the children with the most pronounced electrodermal hyporesponsiveness at age 16 were especially prone to receive an affective disorder diagnosis with the age of 25, which confirms that electrodermal hyporesponsiveness is not specific for developing a schizophrenic disease (Sect. 3.4.1.3).

  238. 238.

    Note that for purposes of transforming SCR amp. into SRR amp. and vice versa, SCLs and SRLs are required (Sect. 2.3.3.2).

  239. 239.

    Data for the two-trial habituation criterion were also analyzed, yielding similar results.

  240. 240.

    Results on habituation speed are largely dependent on the method used (Sect. 3.1.1.3). The trials-to-habituation criterion disregards initial amplitude differences, thus bearing the danger of misclassifying individuals showing a high amplitude to the first stimulus with subsequent borderline but not below-criterion EDRs as slow habituators. Zahn et al. (1968, 1981a) who defined habituation in terms of the EDR amp. decline relative to the trial block with the largest mean EDR amp. found schizophrenics to be slower habituators than normals.

  241. 241.

    At variance with these results are those of Schnur et al. (1989) who found significantly wider third ventricles (by means of computer tomography) in nine schizophrenic responders as compared to 15 nonresponders (SCRs and fingerpulse responses to three 60 dB, 1 kHz tones).

  242. 242.

    In a pilot study with few participants, Hazlett, Dawson, Buchsbaum, and Nuechterlein (1993) found that EDA nonresponders had significantly lower metabolic rates in medial frontal areas, in the hippocampus, and in the amygdala, i.e., in areas that are suggested being involved in generating EDA (Sect. 1.3.4.1).

  243. 243.

    Recorded in a series of 15 tones (1 kHz, 70 dB) with standard methodology, using a time window of 0.5–4 s after stimulus onset (Olbrich & Mussgay, 1987).

  244. 244.

    A possible expansion of experimental possibilities for investigating electrodermal hypo- or hyperresponsivity in schizophrenics as compared to normal controls had been demonstrated by Lim et al. (1999b), using Lim et al.’s (1997) mathematical solution for evaluating overlapping EDRs during short-ISI paradigms, which also offers a possibility to combine EDA with ERP recordings (e.g., Roth, Goodale, & Pfefferbaum, 1991). In comparison with 50 normal controls, their 30 medicated schizophrenics showed a reduced SC-OR response rate, reduced proportion of responders, SCR amp., rise time, peak latency, and steady-state response amplitude during the presentation of 40 tones with ISIs down to about 2.5 s, but no group differences in SCR lat. or SCL emerged.

  245. 245.

    Schell, Dawson, Nuechterlein, Subotnik, and Ventura (2002) investigated the 1-year temporal stability (reliability) of electrodermal measures in 71 young, recent-onset, drug-treated schizophrenics and 36 demographically matched controls (18 and 28% females, respectively) with the methodology used by Dawson et al. (1994). Reliabilities were significant for most EDA variables but generally lower in schizophrenics as in controls, which may be attributable to poorer general arousal regulation (Sect. 3.2.1.1) in patients.

  246. 246.

    It had been discussed in Sect. 3.4.1.3 that depressive patients tend to show the reverse pattern (Venables, 1983). However, Iacono, and Tuason (1983) could not find consistent bilateral asymmetries in EDA in a 1-year follow-up with 26 unipolar and 24 bipolar depressives.

  247. 247.

    Measured as SCRs occurring between 0.8 and 5 s after stimulus onset, exceeding a 1 mm criterion with a maximum gain setting of 0.02 μS/cm.

  248. 248.

    The speech anxiety paradigm used here was a very realistic extension of the one used by Boucsein and Wendt-Suhl (1980). A group of so-called speech experts was presented to the participant on a video screen, which was “joined” by the female experimenter (who could be seen from her back being in front of the group, which required her to always wear the same white collar and hair style throughout the whole study) who presented (over an auditory mixing device) the personal data of the actual participant who prepared the speech.

  249. 249.

    Measured with a 10 Hz and 0.5 V constant voltage system, using dry electrodes made from silver-plated nylon tissue (3.2 cm2 area), taken from palmar finger sites, and monitored telemetrically.

  250. 250.

    Recorded with nonpolarizing electrodes palmar/dorsal from the left hand, amplified by an EEG coupler. The number of SPRs was individually related to productivity, to reduce interindividual differences.

  251. 251.

    Measured with 500 Hz, 1, and 10 kHz with constant voltage (1 V) (SYL transformed to SZL) using 3 × 4 cm V2A nets as electrodes at the backside of the lower legs while working at a punch press, and at the lower arm’s inside during bicycle ergometer task (performed ten times each with 0, 5, 10, 15, and 20 mkp/s, and six times with 25 mkp/s, with 7.5 min duration).

  252. 252.

    The highest amount of EDA was, however, recorded during speaking, which must be regarded as being mainly due to an artifact (Sect. 2.2.5.2).

  253. 253.

    Measured five times successively at 2-s intervals with 10 mA constant current, from the first and second fingers of the left hand, using 1 cm2 aluminium electrodes.

  254. 254.

    Recorded from the left little finger with nonreported methodology.

  255. 255.

    In all studies of the present author’s group, SC was measured with standard methodology.

  256. 256.

    However, caution has to be expressed with artifacts stemming from gross body movements. Goldstein and Shapiro (1988) found that cardiovascular parameters are more sensitive to postural changes during laboratory performance tests (mental arithmetic and isometric handgrip) than EDA. In addition, there was a marked increase in SCL as a transient response during standing up.

  257. 257.

    With stainless steel electrodes, covered with isotonic cream made from Unibase and attached to the right index and the middle finger.

  258. 258.

    Measured with AC from the left hand’s fingers, using a Wheatstone bridge.

  259. 259.

    Recorded with AC (65 Hz), using 10 μA/cm2 average current density, from the participant’s fingers, data transformed into conductance changes.

  260. 260.

    Recorded volar/dorsal from the left foot with 1 cm diameter electrodes filled with a cream from bentonite, glycerine, and Ringer’s salt solution, using a constant current of 5 μA/cm2 and a Wheatstone bridge.

  261. 261.

    Beckman Ag/AgCl electrodes and electrode cream with 0.1N chloride concentration were used to record from the dorsal side of the hand with 12 μA/cm2 constant current. For evaluation, psychophysiological responses were time-shifted 1 s earlier, in to compensate for the time delay associated with EDRs.

  262. 262.

    AC recordings with a 5.25 Hz constant voltage (1 V) source between two Ag/AgCl electrodes, filled with Hellige isotonic electrode cream, attached according to Fig. 2.7 (Sect. 2.2.1.1) to the medial side of the left foot. EDA was recorded on tape with PCM (pulse code modulation) electronics. EDRs were classified into four groups: more than 10, 8–10, 5–7, and 2–4% change with respect to EDL, forming arbitrary units that were averaged per minute.

  263. 263.

    Recorded with standard methodology from the left foot according to Fig. 2.7 with a sampling rate of 20 Hz, evaluated as SCL and as NS.SCR freq. by means of a spline curve for detecting single responses, using an amplitude criterion of 0.05 μS.

  264. 264.

    With Ag/AgCl electrodes, low-pass filtering of 3 and 10 Hz sampling frequency. The amplitude criterion was set to 1 mV.

  265. 265.

    Measured at two sites: At the first and middle finger of the left hand and at the arch of the left foot, with unreported methodology, sampled at 31 Hz.

  266. 266.

    Measured with Beckman Ag/AgCl electrodes and Beckman cream from palmar/dorsal sites with 0.5 V constant voltage. The highest SCR amp. within every 5-s section was evaluated.

  267. 267.

    In fact, the attachment of EDA electrodes being visible for everyone raises problems of compliance in pilots, which may explain some of its infrequent use in that field.

  268. 268.

    Recorded with AC from the ankle, using the methodology of Faber (1983).

  269. 269.

    Ag/AgCl electrodes placed on the arch of the right foot, 3.2 cm apart, filled with Grass EC33 electrode cream. EDA was recorded as SC with the Vitaport 2 ambulatory monitoring device (Sect. 2.2.3.4), together with the other measures.

  270. 270.

    Measured with a custom-made digital recorder with constant voltage across a pair of nickel electrodes attached to the volar surface of the third digit, with the output transformed to SR and recorded as mean SRL for 1 s intervals.

  271. 271.

    Recorded as SC by the Varioport System with standard methodology from the left foot (Fig. 2.7) with a special fixation technique described at the end of Sect. 2.2.2.1.

  272. 272.

    Recorded from the dorsal side of one hand by small suction cap electrodes, using a nonspecified psychogalvanometer.

  273. 273.

    Obtained with standard methodology, 0.01 μS amplitude criterion; overlapping SCRs treated according to evaluation method B, Fig. 2.16, Sect. 2.3.1.2 “Amplitudes of Exosomatic Responses Recorded with Direct Current”.

  274. 274.

    Measured as SR with “a small electric current of constant intensity” from nonspecified palmar electrodes.

  275. 275.

    With Ag/AgCl electrodes filled with 0.5% NaCl electrode cream and 0.4 V constant voltage from the nondominant hand’s palmar sites.

  276. 276.

    In this and the following studies, EDA was recorded with standard methodology.

  277. 277.

    The US Congress enacted the 1988 Employee Polygraph Protection Act (EPPA) to curtail among other things, abuses reported as a result of the widespread use of the polygraph. Problems observed prior to EPPA included poorly standardized and unregulated field practices, and inadequately standardized training for field practitioners, and included cost-cutting and other competitive marketing efforts that led to the proliferation of “chart-rolling” practices which included the conduct of numerous short and unreliable examinations. Foremost among those problems was the selection of examination targets with unproven contribution to the desired outcomes of employee training success and employee integrity. Despite the restriction imposed by EPPA, there are remaining provisions that allow for government and public safety pre-employment polygraph screening, in addition to potential screening for employees in pharmaceutical and nuclear energy industries.

  278. 278.

    A comparison of the estimated number of polygraph examiners was provided by Barland (1988, Table 7.1).

  279. 279.

    The author thanks Mark Handler for his contributions to this section, especially to its polygraph-related part.

  280. 280.

    The abbreviation CQT was introduced initially for “Control Question Test.” However, the term “control” has been replaced by “comparison” in modern polygraph use, since the term control as used in the original CQT did not meet its standard scientific sense. Control questions could only be accepted as an experimental control if they were comparable to the critical questions in all respects except for the process of deception that is under investigation (for more details, see Furedy & Ben-Shakhar, 1991).

  281. 281.

    Despite this putative superiority of the GKT over the CQT, the practical problem arises that the GKT requires the details of the crime to be used in the test having to be kept secret in the public, which may run counter to normal police procedures; this is regarded as a major reason for its infrequent use by polygraphers (Furedy & Heslegrave, 1988).

  282. 282.

    Taken from an U.S. Congress Office of Technology Assessment Report in 1983. Steller (1987) summarized results from laboratory studies, 11 performed with the CQT, and seven with the GKT. According to these, the CQT correctly classified 15.8–90% of the innocent examinees, with false positives ranging from 4.2 to 31.6%. On the other hand, five of the GKT studies made 100% correct classifications, the remaining two studies yielding 88% correct innocents and 12% false positives. With respect to correct classifications of guilty examinees, both methods did not differ considerably (between 60 and 100% correct classifications).

  283. 283.

    In pre- or postemployment screening situations, a slightly different question format is employed, labeled the Relevant Control Test by Lykken (1981).

  284. 284.

    In addition to various proofs of EDA being the most sensitive measure for the detection of deception in comparison with others, as found for example by Dawson (1980) in his study described below (comparing EDA with cardiovascular and respiratory measures), a comparison of an univariate with a multivariate statistical evaluation of CQT data performed by Kircher and Raskin (1988) also yielded a superiority of SCR amp. over blood pressure, respiratory, and vasomotor responses with both methods of evaluation.

  285. 285.

    The ease of detecting psychophysiological correlates of emotion-relevant thoughts by phasic EDA is also used in obscure religious practice, where then so-called E-meters (simple SR monitoring devices) are used as a “scientific” tool to uncover hidden information in participants.

  286. 286.

    The GKT mock-crime paradigm has been used quite a lot by Ben-Shakar and his group for making stimuli significant in their OR studies (Sect. 3.1.1.1). In some cases, the experimental group only imagined having committed a mock-crime, the details of which are given to the participants by instruction, which constitutes a weaker experimental manipulation. The details given to the control group serve the purpose of making the circumstances of the mock crime known to them, to restrict group differences to having committed the mock crime or not.

  287. 287.

    On the other hand, validity criteria of being guilty or innocent are not easy to obtain in field studies, since in the U.S. justice practice false admissions of guilt are common to reduce the amount of penalty.

  288. 288.

    Such a time window for EDR evaluation was recommended by Raskin (1979, p. 597).

  289. 289.

    Raskin (1979) recommended that the EDR amp. should be expressed in terms of mm of chart deflection, which had been regarded by him as producing more reliable results as compared to changes in μS.

  290. 290.

    Recorded with Beckman Ag/AgCl electrodes thenar/hypothenar from the right hand with a current density of 3.8 μA/cm2, using AC coupling with 0.3 s time constant.

  291. 291.

    Recorded with 2.5 × 2.5 cm stainless steel electrodes from the palmar fingertips of the left hand, using 10 μA constant current. For the CQT form with immediate response, the largest decrease in SR which occurred between 1 s following the question onset and 5 s following the verbal answer was evaluated. For the delayed response CQT form, a time window between 1 s following the question onset and 1 s following completion of the question was used to obtain the question SRR, while the largest deflection within 1 and 5 s following the verbal answer was taken as the answer SRR.

  292. 292.

    Recorded with 0.5 V and Ag/AgCl electrodes filled with K–Y jelly, attached to thenar/hypothenar sites of the left hand. The maximum SCR (minimal amplitude criterion 0.05 μS) between 1 and 5 s after stimulus onset was scored and square-root transformed.

  293. 293.

    Recorded as SRL with 10 mA constant current using Beckman Ag/AgCl electrodes from palmar finger sites, converted into SC units. SCL data were reported from 2 s after the question (where the participant answered “no”) to 14 s after the question.

  294. 294.

    Recorded with standard methodology, however, using K–Y jelly and 0.74 V constant voltage. The amplitude criterion was 0.05 μS. For detection of deception, the amplitude criterion was lowered to 0.025 μS. The time windows started 1.5 s, ending 9 s for the CQT and 5 s for the GKT, after stimulus onset.

  295. 295.

    Recorded as the highest deflection from prestimulus level with standard methodology (using Beckman NaCl cream) from the palmar finger sites of the left hand (previously cleaned with soap and water), within 1–5 s following question onset. SCR amp. were expressed in mm (using 0.5 mm as amplitude criterion) and converted to μS response magnitudes.

  296. 296.

    Recorded with 4 kHz AC using lead electrodes.

  297. 297.

    Measured with liquid electrolytes made from polyethylene glycol with 0.9% NaCl (Sect. 2.2.6.3), the active electrode made from platinum, and the reference electrode made from Ag/AgCl. A frequency range from 5 Hz to 500 kHz was used.

  298. 298.

    EDA was recorded in the laboratory with 100 Hz constant current less than 8 μA, using 0.5 cm2 gold cup electrodes with standard NaCl Unibase cream, from alcohol-scrubbed left-hand sites. In the field part of the study, finger tip electrodes which consisted of silver fibers applied to Velcro pads were used for EDA biofeedback. SZLs were transformed to log SYLs for evaluation.

  299. 299.

    A single case study of successful EDA biofeedback was published by Moan (1979), training a 28-year-old female urticaria patient 8 weeks with acoustic/optic EDA biofeedback combined with a relaxation training. Her mean SCL could be reduced from 12 to 7 μS (during the relaxation even to 4 μS), and the skin disease disappeared. There was no relapse seen in an 8 month follow-up.

  300. 300.

    Klaschka (1979) reported an increase in the mitotic index by a factor of 20, 2 days after 30 cellophane-tape stripping, returning slowly to baseline from the third day on.

  301. 301.

    Recorded with 1.6 kHz and 1 mA current density, using metal electrodes of 6 mm diameter, and a 5 mm diameter filter paper soaked with 0.9% NaCl solution, from the inner aspect of the forearm.

  302. 302.

    Measured with a liquid electrode made from a syringe, combining Ag/AgCl with physiological saline in sodium methylcellulose; the inactive electrode placed in the mouth.

  303. 303.

    Measured with 1.6 kHz, using 3 mm steel electrodes together with 4 mm diameter filter paper, soaked with 0.9% NaCl solution.

  304. 304.

    From isopropyl alcohol cleaned sites. The electrolyte for SP consisted of 0.05 M NaCl in a mixture of 50% Unibase and 50% polyethylene glycol, to keep the stratum corneum minimally hydrated (Sect. 2.2.6.1).

  305. 305.

    Recorded with 3 mm2 electrodes from skin washed with soap and water, followed by rubbing with alcohol.

  306. 306.

    Recorded with lead-coated stainless steel electrodes by the use of a digital ohmmeter.

  307. 307.

    With 10 mm diameter stainless steel electrodes and commercial electrode cream from palmar vs. dorsal hand as well as foot sites, from anterior vs. posterior surfaces of the upper arm, and from patella vs. popliteal fossa sites.

  308. 308.

    With gold cup EMG disc electrodes from the ventral and dorsal surface of the right hand, using a bandpass filter of 0.3 Hz–3 kHz.

  309. 309.

    Recorded with standard methodology, using Hellige isotonic electrode cream; time window 1–5 s after stimulus onset; amplitude criterion 500 Ω.

  310. 310.

    Recorded with 9 mm diameter Ag/AgCl electrodes filled with commercial electrode cream from palmar/dorsal sites previously washed with acetone. When laterality of the damage was known, recording was performed at the ipsilateral hand.

  311. 311.

    There were also attempts using NS.EDR freq. and areas under the curve (Sect. 2.3.1.4) to predict awakening from total intravenous anesthesia. However, these parameters had considerably lower predictabilities compared to the “bispectral index”; based on a complex EEG evaluation (Ledowski et al., 2007).

  312. 312.

    Recorded with Ag/AgCl electrodes from the hypothenar eminence or from the palmar side of the little finger as active electrode, with the dorsal forearm as inactive and the ventral forearm as ground electrode. SPR amp. was detected as peak-to-trough measure between 1 and 5 s after stimulus presentation.

  313. 313.

    Recorded with 10 mm2 gold-plated electrodes from palmar finger sites. SCR amp. was calculated as the square-root of the difference between previous SCL and the maximum SCL within 5 s following stimulus onset.

  314. 314.

    Based on the results of an experiment with 20 healthy participants, Tranel, Fowles, and Damasio (1985) advocated the usability of EDRs in the detection of the ability to discriminate between familiar and unfamiliar faces of prosopagnosic patients, who have lost the ability to recognize faces.

  315. 315.

    Recorded with 0.9 cm diameter Ag electrodes palmar/dorsal simultaneously from both hands and feet.

  316. 316.

    Recorded from palmar and plantar surfaces simultaneously with indifferent electrodes attached to the volar side of the forearm (Fig. 2.6) and the lower part of the shinbone (Fig. 2.7), respectively, with otherwise unreported methodology.

  317. 317.

    Recorded as SR with 10 μA constant current unipolarly (thumb vs. upper arm), using 15 mm diameter Ag/AgCl electrodes, transformed to log SC values. Amplitude criterion for NS.SCRs = 0.003 log μS.

  318. 318.

    Recorded with 10 μA/cm2 CC via Ag/AgCl electrodes from the thenar region of the hand. The skin was abraded with sand paper and cleaned with alcohol before electrode attachment.

  319. 319.

    Recorded bilaterally as SCRs with standard methodology, but using 0.5 KCl-agar cream. Time window: 3 s after stimulus onset; amplitude criterion = 0.05 μS.

  320. 320.

    Recorded with standard methodology, however, using hypertonic cream. Time window: 1–5 s after stimulus onset. Habituation criterion: three consecutive SRRs below 500 Ω (Sect. 3.1.1.3).

  321. 321.

    Recorded bilaterally with standard methodology during three series of acoustic stimuli (eight 85 dB tones, twelve 105 dB tones, and two familiar sounds).

  322. 322.

    From the hand and foot opposite to stimulation by gold-plated disc electrodes, using the tip of the fifth finger and the thenar eminence, or the tip of the great toe and the anterior plantar area, respectively, as recording sites.

  323. 323.

    Measured with Ag/AgCl electrodes affixed with Velcro straps to the middle phalanges of the left hand, voltage unreported.

  324. 324.

    Recorded with an active roller electrode without cream and an ECG electrode with hypertonic cream as reference at the leg or arm, by use of a dermohmeter.

  325. 325.

    Janssen, Arntz, and Bouts (1998) used SCL recordings in an investigation into the effect of epinephrine on heat pain threshold in 24 participants (12 of each gender). Focusing of attention exerted a much stronger influence on pain responses than epinephrine infusions.

  326. 326.

    With Ag/AgCl electrodes filled with 0.5% potassium chloride in 2% agar-agar as electrolyte. SCRs were identified as ORs if being greater than 0.05 μS and occurring 1–3 s after tone onset.

  327. 327.

    Recorded from volar finger sites with Ag/AgCl electrodes and otherwise unreported methodology.

  328. 328.

    Recorded with standard methodology, using an MR capable sensor (Brain Products GmbH, Gilching, Germany). After applying MR artifact correction procedures and a bandpass filter of 0.016–5 Hz to the SC signal, SCR amp. was determined as the peak-to-peak difference between the minimum and the maximum SCL within the 1–8 s poststimulus interval.

  329. 329.

    Recorded with 50 μA CC from the first and third fingertip of the right hand through 4 mm diameter Ag/AgCl electrodes with Cambridge electrode cream.

  330. 330.

    Recorded through standard Ag/AgCl electrodes, filled with Parker Labs Signa Gel, by a Grass Physiodata Amplifier System with a DC amplifier.

  331. 331.

    Recorded from the distal phalanges of the first and second fingers of the dominant hand with standard Ag/AgCl electrodes, using physiological NaCl in agar-agar paste as electrolyte. SC was expressed in μS/cm2.

  332. 332.

    The stimulus conditions were listening to classic music and a word association task (five neutral words and five with high emotional content).

  333. 333.

    Recorded with 8 mm diameter Ag/AgCl electrodes filled with isotonic KCl cream from palmar finger sites. The amplitude criterion was 0.05 μS.

  334. 334.

    In another study with normotensives, established and borderline hypertensives (N = 12 in each of the three groups), Fredrikson and Engel (1985) observed higher SCLs in borderline cases compared with the other groups. SC results paralleled cardiovascular results only in normotensives and did not yield hyperreactivity in hypertensives as did HR.

  335. 335.

    EDA has also been applied in parapsychology – an area which is clearly outside scientific research. In their review of the hitherto published 25 investigations using EDA in “direct mental interaction with living systems” or “remote staring,” Schmidt and Walach (2000) came to the conclusion that not even one of the studies referred to the standards mentioned at the beginning of the third chapter of the present book. Therefore, they highly recommended further experimenters in this area to adhere to these standards. Maybe parapsychology is not the most suitable candidate for demonstrating the negative effects of methodological pitfalls, but adhering to standard methodology as recommended here can be regarded as a minimum requirement in all kinds of applied electrodermal research. Stevens (2000) reanalyzed SC data from two previously conducted studies with altogether 123 participants, collected during periods in which a nonpresent “sender” tried to activate or calm the participants. A resting condition served as control. For better interindividual comparability, SC data were transformed to standardized z-scores according to (2.20a) in Sect. 2.3.3.4 “Transformation into Standard Values”. Several individuals showed an occasional increase in SC shortly after a presumed stimulus onset, i.e., the start of a period during which they were assumed being mentally influenced by the “sender,” but no clear pattern emerged that could have allowed a distinction between these changes and spontaneous EDRs. Some of the changes were also in the opposite direction as intended. Interestingly, SC emerged significantly more variability during the two “influence” periods as compared to the resting condition. Schmidt, Schneider, Binder, Bürkle, and Walach (2001) analyzed SC data (recorded with standard methodology from thenar/hypothenar eminences) from 26 sessions with a total of 52 participants (24 females, 28 males). The room temperature was set to either 23 or 26°C to enhance electrodermal reactivity (Sect. 2.4.1.1). In addition to SC, respiration was recorded, not so much for the control of artifacts but as a mediator for explaining some of the variance in phasic EDA (Sect. 2.2.5.2). Each session consisted of ten activate and ten calm epochs. The “sender” viewed the other participant’s SC curve, trying to activate or deactivate the participant by means of mentality or intentionality. The randomized influence epochs were interspersed by 15-s resting periods. The authors applied five different procedures for evaluating their data, four of which yielded approximately the same results. Besides the SCL, NS.SCR freq., and mean NS.SCR amp. were evaluated. All three parameters showed a marked effect of the “activate” condition. However, since EDA substantially covaried with respiratory activity, the authors interpreted their results as demonstrating a sort of global effect of their “mental interaction” condition on the psychophysiological arousal of their participants.

References

  • Aberg, P., Geladi, P., Nicander, I., Hansson, J., Holmgren, U., & Ollmar, S. (2005). Non-invasive and microinvasive electrical impedance spectra of skin cancer – A comparison between two techniques. Skin Research and Technology, 11, 281–286.

    Article  PubMed  Google Scholar 

  • Akiskal, H. S., & McKinney, W. T. (1975). Overview of recent research in depression. Archives of General Psychiatry, 32, 285–305.

    PubMed  Google Scholar 

  • Alexander, G. E., Crutcher, M. D., & DeLong, M. R. (1990). Basal ganglia-thalamocortical circuits: Parallel substrates for motor, oculomotor, “prefrontal” and “limbic” functions. Progress in Brain Research, 85, 119–146.

    Article  PubMed  Google Scholar 

  • American Psychiatric Association. (2000). Diagnostic and statistical manual of mental disorders (4th ed. revised (DSM IV-R)). Washington, DC: APA.

    Google Scholar 

  • Andersson, S., & Finset, A. (1998). Heart rate and skin conductance reactivity to brief psychological stress in brain-injured patients. Journal of Psychosomatic Research, 44, 645–656.

    Article  PubMed  Google Scholar 

  • Andersson, S., & Finset, A. (1999). Electrodermal responsiveness and negative symptoms in brain injured patients. Journal of Psychophysiology, 13, 109–116.

    Article  Google Scholar 

  • Andersson, S., Gundersen, P. M., & Finset, A. (1999). Emotional activation during therapeutic interaction in traumatic brain injury: Effect of apathy, self-awareness and implications for rehabilitation. Brain Injury, 13, 393–404.

    Article  PubMed  Google Scholar 

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

    Google Scholar 

  • Annett, M. (1982). Handedness. In J. G. Beaumont (Ed.), Divided visual field studies of cerebral organization (pp. 195–215). New York: Academic.

    Google Scholar 

  • Arnett, P. A. (1997). Autonomic responsivity in psychopaths: A critical review and theoretical proposal. Clinical Psychology Review, 17, 903–936.

    Article  PubMed  Google Scholar 

  • Arnett, P. A., & Newman, J. P. (2000). Gray’s three-arousal model: An empirical investigation. Personality and Individual Differences, 28, 1171–1189.

    Article  Google Scholar 

  • Ax, A. F. (1953). The physiological differentiation between fear and anger in humans. Psychosomatic Medicine, 15, 433–442.

    PubMed  Google Scholar 

  • Ax, A. F., & Bamford, J. L. (1970). The GSR recovery limb in chronic schizophrenics. Psychophysiology, 7, 145–147.

    Article  PubMed  Google Scholar 

  • Backs, R. W., & Boucsein, W. (2009). Psychophysiology in digital human modeling. In V. G. Duffy (Ed.), Handbook of digital human modeling (pp. 16–1–16–14). Boca Raton: CRC.

    Google Scholar 

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

    Article  Google Scholar 

  • Baltissen, R., & Boucsein, W. (1986). Effects of a warning signal to aversive white noise stimulation: Does warning “short-circuit” habituation? Psychophysiology, 23, 224–231.

    Article  PubMed  Google Scholar 

  • Baltissen, R., & Weimann, C. (1989). Orienting reaction reinstatement or preception? Effects of predictability on reactions to pink noise stimulation of different intensities. Psychophysiology, 26, S12.

    Google Scholar 

  • Bankart, C. P., & Elliot, R. (1974). Heart rate and skin conductance in anticipation of shocks with varying probability of occurrence. Psychophysiology, 11, 160–174.

    Article  PubMed  Google Scholar 

  • Freixa i Baqué, E., Catteau, M.-C., Miossec, Y., & Roy, J.-C. (1984). Asymmetry of electrodermal activity: A review. Biological Psychology, 18, 219–239.

    Google Scholar 

  • Freixa i Baqué, E., Chevalier, B., Grubar, J. C., Lambert, C., Lancry, A., Leconte, P., Meriaux, H., & Spreux, F. (1983). Spontaneous electrodermal activity during sleep in man: An intranight study. Sleep, 6, 77–81.

    Google Scholar 

  • Freixa i Baqué, E., & de Bonis, M. (1983). Electrodermal asymmetry during human sleep. Biological Psychology, 17, 145–151.

    Google Scholar 

  • Barland, G. H. (1988). The polygraph in practice. In A. Gale (Ed.), The polygraph test: Lies, truth and science (pp. 73–95). London: Sage.

    Google Scholar 

  • Barland, G. H. (1999). American Polygraph Association Newsletter, 32, 16–17.

    Google Scholar 

  • Barland, G. H., & Raskin, D. C. (1973). Detection of deception. In W. F. Prokasy & D. C. Raskin (Eds.), Electrodermal activity in psychological research (pp. 417–477). New York: Academic.

    Google Scholar 

  • Barry, R. J. (1975). Low-intensity auditory stimulation and the GSR orienting response. Physiological Psychology, 3, 98–100.

    Google Scholar 

  • Barry, R. J. (1976). Failure to find the “local” EEG OR to low-level auditory stimulation. Physiological Psychology, 4, 171–174.

    Google Scholar 

  • Barry, R. J. (1982). Novelty and significance effects in the fractionation of phasic OR measures: A synthesis with traditional OR theory. Psychophysiology, 19, 28–35.

    Article  PubMed  Google Scholar 

  • Barry, R. J. (1987). Preliminary process in orienting response elicitation. In P. K. Ackles, J. R. Jennings, & M. G. H. Coles (Eds.), Advances in Psychophysiology (Vol. 2, pp. 131–195). Greenwich, CT: Jai.

    Google Scholar 

  • Barry, R. J. (1990). Scoring criteria for response latency and habituation in electrodermal research: A study in the context of the orienting response. Psychophysiology, 27, 94–100.

    Article  PubMed  Google Scholar 

  • Barry, R. J. (2004). Stimulus significance effects in habituation of the phasic and tonic orienting reflex. Integrative Physiological and Behavioral Science, 39, 166–179.

    Article  PubMed  Google Scholar 

  • Barry, R. J., Clarke, A. R., Johnstone, S. J., & Rushby, J. A. (2008). Timing of caffeine’s impact on autonomic and central nervous system measures: Clarification of arousal effects. Biological Psychology, 77, 304–316.

    Article  PubMed  Google Scholar 

  • Barry, R. J., & O’Gorman, J. G. (1987). Stimulus omission and the orienting response: Latency differences suggest different mechanisms. Biological Psychology, 25, 261–276.

    Article  PubMed  Google Scholar 

  • Barry, R. J., & Sokolov, E. N. (1993). Habituation of phasic and tonic components of the orienting reflex. International Journal of Psychophysiology, 15, 39–42.

    Article  PubMed  Google Scholar 

  • Bartfai, A., Edman, G., Levander, S. E., Schalling, D., & Sedvall, G. (1984). Bilateral skin conductance activity, clinical symptoms and CSF monoamine metabolite levels in unmedicated schizophrenics, differing in rate of habituation. Biological Psychology, 18, 201–218.

    Article  PubMed  Google Scholar 

  • Bauer, R. M. (1986). The cognitive psychophysiology of prosopagnosia. In H. Ellis, M. Jeeves, F. Newcombe, & A. Young (Eds.), Aspects of face processing (pp. 253–267). Dordrecht: Nijhoff.

    Google Scholar 

  • Beatty, J. (1983). Biofeedback in theory and practice. In A. Gale & J. A. Edwards (Eds.), Physiological correlates of human behaviour (Individual differences and psychopathology, Vol. 3, pp. 233–246). London: Academic.

    Google Scholar 

  • Beatty, J., & Legewie, H. (Eds.). (1977). Biofeedback and behavior. New York: Plenum.

    Google Scholar 

  • Beauchaine, T. P., Katkin, E. S., Strassberg, Z., & Snarr, J. (2001). Disinhibitory psychopathology in male adolescents: Discriminating conduct disorder from attention-deficit/hyperactivity disorder through concurrent assessment of multiple autonomic states. Journal of Abnormal Psychology, 110, 610–624.

    Article  PubMed  Google Scholar 

  • Beaumont, J. G. (Ed.). (1982). Divided visual field studies of cerebral organization. New York: Academic.

    Google Scholar 

  • Bechara, A., & Damasio, H. (2002). Decision-making and addiction (part I): Impaired activation of somatic states in substance dependent individuals when pondering decisions with negative future consequences. Neuropsychologia, 40, 1675–1689.

    Article  PubMed  Google Scholar 

  • Bechara, A., & Damasio, A. R. (2005). The somatic marker hypothesis: A neural theory of economic decision. Games and Economic Behavior, 52, 336–372.

    Article  Google Scholar 

  • Bechara, A., Damasio, A. R., Damasio, H., & Anderson, S. W. (1994). Insensitivity to future consequences following damage to human prefrontal cortex. Cognition, 50, 7–15.

    Article  PubMed  Google Scholar 

  • Bechara, A., Damasio, H., Damasio, A. R., & Lee, G. P. (1999). Different contributions of the human amygdala and ventromedial prefrontal cortex to decision-making. The Journal of Neuroscience, 19, 5473–5481.

    PubMed  Google Scholar 

  • Bechara, A., Damasio, H., Tranel, D., & Damasio, A. R. (2005). The Iowa Gambling Task and the somatic marker hypothesis: Some questions and answers. Trends in Cognitive Sciences, 9, 159–162.

    Article  PubMed  Google Scholar 

  • Bechara, A., Dolan, S., Denburg, N., Hindes, A., Anderson, S. W., & Nathan, P. E. (2001). Decision making deficits, linked to a dysfunctional ventromedial prefrontal cortex, revealed in alcohol and stimulant abusers. Neuropsychologia, 39, 376–389.

    Article  PubMed  Google Scholar 

  • Bechara, A., Dolan, S., & Hindes, A. (2002). Decision-making and addiction (part II): Myopia for the future or hypersensitivity to reward? Neuropsychologia, 40, 1690–1705.

    Article  PubMed  Google Scholar 

  • Bechara, A., Tranel, D., & Damasio, H. (2000). Characterization of the decision-making deficit of patients with ventromedial prefrontal cortex lesion. Brain, 123, 2189–2202.

    Article  PubMed  Google Scholar 

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

    Article  PubMed  Google Scholar 

  • Ben-Shakhar, G. (1980). Habituation of the orienting response to complex sequences of stimuli. Psychophysiology, 17, 524–534.

    Article  PubMed  Google Scholar 

  • Ben-Shakhar, G. (1994). The roles of stimulus novelty and significance in determining the electrodermal orienting response: Interactive vs. additive approaches. Psychophysiology, 31, 402–411.

    Article  PubMed  Google Scholar 

  • Ben-Shakhar, G., Asher, T., Poznansky-Levy, A., Asherowitz, R., & Lieblich, I. (1989). Stimulus novelty and significance as determinants of electrodermal responsivity: The serial position effect. Psychophysiology, 26, 29–38.

    Article  PubMed  Google Scholar 

  • Ben-Shakhar, G., & Dolev, K. (1996). Psychophysiological detection through the guilty knowledge technique: The effects of mental countermeasures. Journal of Applied Psychology, 81, 273–281.

    Article  PubMed  Google Scholar 

  • Ben-Shakhar, G., & Elaad, E. (2003). The validity of psychophysiological detection of information with the Guilty Knowledge Test: A meta-analytic review. Journal of Applied Psychology, 88, 131–151.

    Article  PubMed  Google Scholar 

  • Ben-Shakhar, G., & Furedy, J. J. (1990). Theories and applications in the detection of deception: A psychophysiological and international perspective. New York: Springer.

    Book  Google Scholar 

  • Ben-Shakhar, G., & Gati, I. (2003). The effects of serial position and frequency of presentation of common stimulus features on orienting response reinstatement. Psychophysiology, 40, 139–145.

    Article  PubMed  Google Scholar 

  • Ben-Shakhar, G., Gati, I., & Salamon, N. (1995). Generalization of the orienting response to significant stimuli: The roles of common and distinctive stimulus components. Psychophysiology, 32, 36–42.

    Article  PubMed  Google Scholar 

  • Ben-Shakhar, G., & Lieblich, I. (1982). The dichotomization theory for differential autonomic responsivity reconsidered. Psychophysiology, 12, 277–281.

    Article  Google Scholar 

  • Ben-Shakhar, G., Lieblich, I., & Kugelmass, S. (1982). Interactive effects of stimulus probability and significance on the skin conductance response. Psychophysiology, 19, 112–114.

    Article  PubMed  Google Scholar 

  • Berlyne, D. E. (1961). Conflict and the orientation reaction. Journal of Experimental Psychology, 62, 476–483.

    Article  PubMed  Google Scholar 

  • Bernat, E., Patrick, C. J., Benning, S. D., & Tellegen, A. (2006). Effects of picture content and intensity on affective physiological response. Psychophysiology, 43, 93–103.

    Article  PubMed  Google Scholar 

  • Bernstein, A. S. (1979). The orienting response as novelty and significance detector: Reply to O’Gorman. Psychophysiology, 16, 263–273.

    Article  PubMed  Google Scholar 

  • Bernstein, A. S., Frith, C. D., Gruzelier, J. H., Patterson, T., Straube, E., Venables, P. H., et al. (1982). An analysis of the skin conductance orienting response in samples of American, British, and German schizophrenics. Biological Psychology, 14, 155–211.

    Article  PubMed  Google Scholar 

  • Bernstein, A. S., Riedel, J. A., Graae, F., Seidman, D., Steele, H., Connolly, J., et al. (1988). Schizophrenia is associated with altered orienting activity; depression with electrodermal (cholinergic?) deficit and normal orienting response. Journal of Abnormal Psychology, 97, 3–12.

    Article  PubMed  Google Scholar 

  • Bernstein, A. S., Schneider, S. J., Juni, S., Pope, A. T., & Starkey, P. (1980). The effect of stimulus significance on the electrodermal response in chronic schizophrenia. Journal of Abnormal Psychology, 89, 93–97.

    Article  PubMed  Google Scholar 

  • Bernstein, A. S., Taylor, K. W., Starkey, P., Juni, S., Lubowski, J., & Paley, H. (1981). Bilateral skin conductance, finger pulse volume, and EEG orienting response to tones of differing intensities in chronic schizophrenics and controls. The Journal of Nervous and Mental Disease, 169, 513–528.

    Article  PubMed  Google Scholar 

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

    Article  PubMed  Google Scholar 

  • Birk, L., Crider, A., Shapiro, D., & Tursky, B. (1966). Operant electrodermal conditioning under partial curarization. Journal of Comparative and Physiological Psychology, 62, 165–166.

    Article  PubMed  Google Scholar 

  • Birket-Smith, M., Hasle, N., & Jensen, H. H. (1993). Electrodermal activity in anxiety disorders. Acta Psychiatrica Scandinavica, 88, 350–355.

    Article  PubMed  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 

  • Bjornaes, H., Smith-Meyer, H., Valen, H., Kristiansen, K., & Ursin, H. (1977). Plasticity and reactivity in unconscious patients. Neuropsychologia, 15, 451–455.

    Article  PubMed  Google Scholar 

  • Blackburn, R. (1983). Psychopathy, delinquency and crime. In A. Gale & J. A. Edwards (Eds.), Physiological correlates of human behaviour (Individual differences and psychopathology, Vol. 3, pp. 187–205). London: Academic.

    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 

  • Bob, P., Susta, M., Glaslova, K., Pavlat, J., & Raboch, J. (2007). Lateralized electrodermal dysfunction and complexity in patients with schizophrenia and depression. Neuroendocrinology Letters, 28, 11–15.

    PubMed  Google Scholar 

  • Böhmelt, A. H., Vanman, E. J., Dawson, M. E., & Boucsein, W. (1983). Modification of the acoustic startle eyeblink reflex with complex visual prestimuli: Late effects of attention. Journal of Psychophysiology, 9, 261–262.

    Google Scholar 

  • Bond, A. J., James, D. C., & Lader, M. H. (1974). Physiological and psychological measures in anxious patients. Psychological Medicine, 4, 364–373.

    Article  PubMed  Google Scholar 

  • Borkovec, T. D. (1970). Autonomic reactivity to sensory stimulation in psychopathic, neurotic, and normal juvenile delinquents. Journal of Consulting and Clinical Psychology, 35, 217–222.

    Article  PubMed  Google Scholar 

  • Boucsein, W. (1973). Analyse einiger psychologischer Testverfahren zur Erfassung von Persönlichkeitsmerkmalen. Unpublished Report of the Psychological Institute, Düsseldorf.

    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. (2000). The use of psychophysiology for evaluating stress-strain processes in human-computer interaction. In R. W. Backs & W. Boucsein (Eds.), Engineering psychology. Issues and applications (pp. 289–309). Mahwah, NY: Lawrence Erlbaum.

    Google Scholar 

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

    Google Scholar 

  • Boucsein, W., & Backs, R. W. (2009). The psychophysiology of emotion, arousal, and personality: Methods and models. In V. G. Duffy (Ed.), Handbook of digital human modeling (pp. 35–1–35–18). 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., & Frye, M. (1974). Physiologische und psychische Wirkungen von Mißerfolgsstress unter Berücksichtigung des Merkmals Repression-Sensitization. Zeitschrift für Experimentelle und Angewandte Psychologie, 21, 339–366.

    PubMed  Google Scholar 

  • Boucsein, W., Greif, S., & Wittekamp, J. (1984). Systemresponsezeiten als Belastungsfaktor bei Bildschirm-Dialogtätigkeiten. Zeitschrift für Arbeitswissenschaft., 38(10 NF), 113–122.

    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., & Ottmann, W. (1996). Psychophysiological stress effects from the combination of night-shift work and noise. Biological Psychology, 42, 301–322.

    Article  PubMed  Google Scholar 

  • Boucsein, W., & Schaefer, F. (2008). Objective emotional assessment of industrial products. In J. Westerink, M. Ouwerkerk, T. Overbeek, F. Pasveer, & B. de Ruyter (Eds.), Probing Experience: From Assessment of User Emotions and Behaviour to Development of Products (pp. 69–76). Dordrecht: Springer.

    Google Scholar 

  • Boucsein, W., Schaefer, F., Kefel, M., Busch, P., & Eisfeld, W. (2002). Objective emotional assessment of tactile hair properties and their modulation by different product worlds. International Journal of Cosmetic Science, 24, 135–150.

    Article  PubMed  Google Scholar 

  • Boucsein, W., Schaefer, F., Schwerdtfeger, A., Busch, P., & Eisfeld, W. (1999). Objective emotional assessment of foam. SÖFW-Journal, 125, 2–17.

    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. (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., Valentin, A., & Furedy, J. J. (1993). Psychophysiological and behavioral differences as a function of age and Parkinson’s disease. The Journal of Integrative Behavioral Science, 28, 213–225.

    Article  Google Scholar 

  • Boucsein, W., & Wendt-Suhl, G. (1976). The effect of chlordiazepoxide on the anticipation of electric shocks. Psychopharmacology, 48, 303–306.

    Article  PubMed  Google Scholar 

  • Boucsein, W., & Wendt-Suhl, G. (1980). An experimental investigation of elements involved in the anticipation of public speaking. Archiv für Psychologie, 133, 149–156.

    PubMed  Google Scholar 

  • Boucsein, W., & Wendt-Suhl, G. (1982). Experimentalpsychologische Untersuchung psychischer und psychophysiologischer Wirkungen von Cloxazolam und Diazepam unter angstinduzierenden und Normalbedingungen bei gesunden Probanden. Pharmacopsychiatria, 15, 48–56.

    Article  PubMed  Google Scholar 

  • Bowman, C. H., & Turnbull, O. H. (2003). Real versus facsimili reinforcers on the Iowa Gambling Task. Brain and Cognition, 53, 207–210.

    Article  PubMed  Google Scholar 

  • Boyd, G. M., & Maltzman, I. (1983). Bilateral asymmetry of skin conductance responses during auditory and visual tasks. Psychophysiology, 20, 196–203.

    Article  PubMed  Google Scholar 

  • Bradley, M. T., & Ainsworth, D. (1984). Alcohol and the psychophysiological detection of deception. Psychophysiology, 21, 63–71.

    Article  PubMed  Google Scholar 

  • Bradley, M. T., & Warfield, J. F. (1984). Innocence, information, and the guilty knowledge test in the detection of deception. Psychophysiology, 21, 683–689.

    Article  PubMed  Google Scholar 

  • Bradshaw, J. L., & Nettleton, N. C. (1981). The nature of hemispheric specialization in man. The Behavioral and Brain Sciences, 4, 51–91.

    Article  Google Scholar 

  • Brand, G., Millot, J. L., Saffaux, M., & Morand-Villeneuve, N. (2002). Lateralization in human nasal chemoreception: Differences in bilateral electrodermal responses related to olfactory and trigeminal stimuli. Behavioural Brain Research, 133, 205–210.

    Article  PubMed  Google Scholar 

  • Braune, S., Albus, M., Fröhler, M., Höhn, T., & Scheibe, G. (1994). Psychophysiological and biochemical changes in patients with panic attacks in a defined situational arousal. European Archives of Psychiatry and Clinical Neuroscience, 244, 86–92.

    Article  PubMed  Google Scholar 

  • Breen, N., Caine, D., & Coltheart, M. (2000). Models of face recognition and delusional misidentification: A critical review. Cognitive Neuropsychology, 17, 55–71.

    Article  PubMed  Google Scholar 

  • Brekke, J. S., Raine, A., & Thomson, C. (1995). Cognitive and psychphysiological correlates of positive, negative and disorganized symptoms in the schizophrenia spectrum. Psychiatry Research, 57, 241–250.

    Article  PubMed  Google Scholar 

  • Brenner, S. L., Beauchaine, T. P., & Sylvers, P. (2005). A comparison of psychophysiological and self-report measures of BAS and BIS activation. Psychophysiology, 42, 108–115.

    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 

  • Broadbent, D. E. (1971). Decision and stress. London: Academic.

    Google Scholar 

  • Broughton, R. J., Poire, R., & Tassinari, C. A. (1965). The electrodermogram (Tarchanoff effect) during sleep. Electroencephalography and Clinical Neurophysiology, 18, 691–708.

    Article  PubMed  Google Scholar 

  • Bryden, M. P. (1965). Tachistoscopic recognition, handedness and cerebral dominance. Neuropsychologia, 3, 1–8.

    Article  Google Scholar 

  • Bryden, M. P. (1979). Evidence for sex related differences in cerebral organization. In M. A. Wittig & A. C. Peterson (Eds.), Sex-related differences in cognitive functioning (pp. 121–143). New York: Academic.

    Google Scholar 

  • Buck, R. (1977). Nonverbal communication of affect in preschool children: Relationships with personality and skin conductance. Journal of Personality and Social Psychology, 35, 225–236.

    Article  PubMed  Google Scholar 

  • Buck, R. (1980). Nonverbal behavior and the theory of emotion: The facial feedback hypothesis. Journal of Personality and Social Psychology, 38, 811–824.

    Article  PubMed  Google Scholar 

  • Buck, R., Miller, R. E., & Caul, W. F. (1974). Sex, personality, and physiological variables in the communication of affect via facial expression. Journal of Personality and Social Psychology, 30, 587–596.

    Article  PubMed  Google Scholar 

  • Buck, R., Savin, V. J., Miller, R. E., & Caul, W. F. (1972). Communication of affect through facial expressions in humans. Journal of Personality and Social Psychology, 23, 362–371.

    Article  PubMed  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 

  • 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 

  • Burch, N. R., & Greiner, T. H. (1960). A bioelectric scale of human alertness: Concurrent recordings of the EEG and GSR. Psychiatric Research Reports of the American Psychological Association, 12, 183–193.

    Google Scholar 

  • Byrne, D. (1961). The repression-sensitization scale: Rationale, reliability, and validity. Journal of Personality, 29, 334–349.

    Article  PubMed  Google Scholar 

  • Cacioppo, J. T., & Petty, R. E. (1986). Social processes. In M. G. H. Coles, E. Donchin, & S. W. Porges (Eds.), Psychophysiology: Systems, processes, and applications (pp. 646–679). Amsterdam: Elsevier.

    Google Scholar 

  • Caffyn, J. M. (1964). Psychological laboratory techniques in copy research. Journal of Advertising Research, 4, 45–50.

    Google Scholar 

  • Cahill, L., & McGaugh, J. L. (1998). Mechanisms of emotional arousal and lasting declarative memory. Trends in Neurosciences, 21, 294–299.

    Article  PubMed  Google Scholar 

  • Cambrai, M., Clar, E. J., Grosshans, E., & Altermatt, C. (1979). Skin impedance and phoreographic index in psoriasis: Relationship with action kinetics of three treatments. Archives of Dermatological Research, 264, 197–211.

    Article  PubMed  Google Scholar 

  • Campbell, M. C., Stout, J. C., & Finn, P. R. (2004). Reduced autonomic responsiveness to gambling task losses in Huntington’s disease. Journal of the International Neuropsychological Society, 10, 239–245.

    Article  PubMed  Google Scholar 

  • Campos, J. J., & Johnson, H. J. (1967). Affect, verbalization, and directional fractionation of autonomic responses. Psychophysiology, 3, 285–290.

    Article  PubMed  Google Scholar 

  • Canavan, A. G. M., Passingham, R. E., Marsden, C. D., Quinn, N., Wyke, M., & Polkey, C. E. (1989). The performance on learning tasks of patients in the early stages of Parkinson’s disease. Neuropsychologica, 27, 141–156.

    Article  Google Scholar 

  • Cannon, T. D., Fuhrmann, M., Mednick, S. A., Machon, R. A., Parnas, J., & Schulsinger, F. (1988). Third ventricle enlargement and reduced electrodermal responsiveness. Psychophysiology, 25, 153–156.

    Article  PubMed  Google Scholar 

  • Cannon, T. D., Mednick, S. A., & Parnas, J. (1989). Genetic and perinatal determinants of structural brain deficits in schizophrenia. Archives of General Psychiatry, 46, 883–889.

    PubMed  Google Scholar 

  • Cannon, T. D., Mednick, S. A., & Parnas, J. (1990). Antecedents of predominantly negative – and predominantly positive-symptom schizophrenia in a high-risk population. Archives of General Psychiatry, 47, 622–632.

    PubMed  Google Scholar 

  • Carbonnell, L., Vidal, F., Sequeira, H., & Caverni, J.-P. (2006). A reasoning bias revealed by electrodermal activity. Psychophysiology, 43, 387–393.

    Article  PubMed  Google Scholar 

  • Carlsson, K., Petersson, K. M., Lundqvist, D., Karlsson, A., Ingvar, M., & Öhman, A. (2004). Fear and the amygdala: Manipulation of awareness generates differential cerebral responses to phobic and fear-relevant (but nonfeared) stimuli. Emotion, 4, 340–353.

    Article  PubMed  Google Scholar 

  • Carney, R. M., Hong, B. A., Kulkarni, S., & Kapila, A. (1981). A comparison of EMG and SCL in normal and depressed subjects. The Pavlovian Journal of Biological Science, 16, 212–216.

    PubMed  Google Scholar 

  • Carrillo, E., Moya-Albiol, L., González-Bono, E., Salvador, A., Ricarte, J., & Gómez-Amor, J. (2001). Gender differences in cardiovascular and electrodermal responses to public speaking task: The role of anxiety and mood states. International Journal of Psychophysiology, 42, 253–264.

    Article  PubMed  Google Scholar 

  • Carstensen, L. L., Issacowitz, D. M., & Charles, S. T. (1999). Taking time seriously: A theory of socioemotional selectivity. The American Psychologist, 54, 165–181.

    Article  PubMed  Google Scholar 

  • Carter, S., & Smith-Pasqualini, M. C. (2004). Stronger autonomic response accompanies better learning: A test of Damasio’s somatic marker hypothesis. Cognition & Emotion, 18, 901–911.

    Article  Google Scholar 

  • Champion, R. A., & Jones, J. E. (1961). Forward, backward, and pseudoconditioning of the GSR. Journal of Experimental Psychology, 62, 58–61.

    Article  PubMed  Google Scholar 

  • Charles, S. T., Reynolds, C. A., & Gatz, M. (2001). Age-related differences and change in positive and negative affect over 23 years. Journal of Personality and Social Psychology, 80, 136–151.

    Article  PubMed  Google Scholar 

  • Chassande, B., Charpentier, G., Budy, I., Willer, J. C., & Lille, F. (1996). Sympathetic skin responses and heart rate variability in diabetic patients with different grades of polyneuropathy. Revue Neurologique, 152, 623–629.

    PubMed  Google Scholar 

  • Chattopadhyay, P. K. (1981). Bilateral skin resistance responses in anxiety. Indian Journal of Clinical Psychology, 8, 29–34.

    Google Scholar 

  • Chattopadhyay, P. K., & Biswas, P. K. (1983). Characteristics of galvanic skin response in anxious patients and normal subjects. Indian Journal of Clinical Psychology, 10, 159–164.

    Google Scholar 

  • Chattopadhyay, P. K., Bond, A. J., & Lader, M. H. (1975). Characteristics of galvanic skin response in anxiety states. Journal of Psychiatric Research, 12, 265–270.

    Article  Google Scholar 

  • Chattopadhyay, P. K., Cooke, E., Toone, B., & Lader, M. (1980). Habituation of physiological responses in anxiety. Biological Psychiatry, 15, 711–721.

    PubMed  Google Scholar 

  • Chattopadhyay, P. K., Mazumdar, P., & Basu, A. K. (1982). Habituation of electrodermal responses in tension-headache sufferers and non-tension headache controls. Indian Journal of Psychiatry, 24, 61–65.

    Google Scholar 

  • Checkley, H. (1964). The mask of sanity. St. Louis: Mosby.

    Google Scholar 

  • Chiu, Y.-C., Lin, C.-H., Huang, J.-T., Lin, S., Lee, P.-L., & Hsieh, J.-C. (2008). Immediate gain is long-term loss: Are there foresighted decision makers in the Iowa Gambling Task? Behavioral and Brain Functions, 4, 4–13.

    Article  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 

  • Claridge, G. S. (1967). Personality and arousal. Oxford: Pergamon.

    Google Scholar 

  • Cleveland, D. E. (1961). Driver tension and rural intersection illumination. Traffic Engineering, 32, 11–16.

    Google Scholar 

  • Coan, J. A., & Allen, J. J. B. (2003). Frontal EEG asymmetry and the behavioural activation and inhibition systems. Psychophysiology, 40, 106–114.

    Article  PubMed  Google Scholar 

  • Coan, J. A., & Allen, J. J. B. (2004). Frontal EEG asymmetry as a moderator and mediator of emotion. Biological Psychology, 67, 7–49.

    Article  PubMed  Google Scholar 

  • Coan, J. A., Allen, J. J. B., & Harmon-Jones, E. (2001). Voluntary facial expression and hemispheric asymmetry over the frontal cortex. Psychophysiology, 38, 912–925.

    Article  PubMed  Google Scholar 

  • Colbert, A. P., Hammerschlag, R., Aickin, M., & McNames, J. (2004). Reliability of the Prognos electrodermal device for measurement of electric skin resistance at acupuncture points. Journal of Alternative and Complementary Medicine, 10, 610–616.

    Google Scholar 

  • Colby, C. Z., Lanzetta, J. T., & Kleck, R. E. (1977). Effects of the expression of pain on autonomic and pain tolerance responses to subject-controlled pain. Psychophysiology, 14, 537–540.

    Article  PubMed  Google Scholar 

  • Coles, M. G. H., & Gale, A. (1971). Physiological reactivity as a predictor of performance in a vigilance task. Psychophysiology, 8, 594–599.

    Article  PubMed  Google Scholar 

  • Coles, M. G. H., Gale, A., & Kline, P. (1971). Personality and habituation of the orienting reaction: Tonic and response measures of electrodermal activity. Psychophysiology, 8, 54–63.

    Article  PubMed  Google Scholar 

  • Colombetti, G. (2008). The somatic marker hypotheses, and what the Iowa Gambling Task does and does not show. The British Journal for the Philosophy of Sciences, 59, 51–71.

    Article  Google Scholar 

  • Correa, E. J., & Adams, H. E. (1981). The validity of the pre-employment polygraph examination and the effects of motivation. Polygraph, 10, 143–155.

    Google Scholar 

  • Corteen, R. S. (1969). Skin conductance changes and word recall. British Journal of Psychology, 60, 81–84.

    Article  PubMed  Google Scholar 

  • Corteen, R. S., & Dunn, D. (1974). Shock-associated words in a nonattended message: A test for momentary awareness. Journal of Experimental Psychology, 102, 1143–1144.

    Article  Google Scholar 

  • Corteen, R. S., & Wood, B. (1972). Autonomic responses to shock-associated words in an unattended channel. Journal of Experimental Psychology, 94, 308–313.

    Article  PubMed  Google Scholar 

  • Crider, A. (1993). Electrodermal response lability-stability: Individual difference correlates. In J.-C. Roy, W. Boucsein, D. C. Fowles, & J. H. Gruzelier (Eds.), Progress in electrodermal research (pp. 173–186). London: Plenum.

    Chapter  Google Scholar 

  • Crider, A. (2008). Personality and electrodermal response lability: An interpretation. Applied Psychophysiology and Biofeedback, 33, 141–148.

    Article  PubMed  Google Scholar 

  • Crider, A., & Augenbraun, C. B. (1975). Auditory vigilance correlates of electrodermal response habituation speed. Psychophysiology, 12, 36–40.

    Article  PubMed  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 

  • Critchley, H. D., Melmed, R. N., Featherstone, E., Mathias, C. J., & Dolan, R. J. (2001). Brain activity during biofeedback relaxation. A functional neuroimaging investigation. Brain, 124, 1003–1012.

    Google Scholar 

  • Critchley, H. D., Melmed, R. N., Featherstone, E., Mathias, C. J., & Dolan, R. J. (2002). Volitional control of autonomic arousal: A functional magnetic resonance study. NeuroImage, 16, 909–919.

    Article  PubMed  Google Scholar 

  • Crone, E. A., Somsen, R. J., van Beek, B., & van der Molen, M. W. (2004). Heart rate and skin conductance analysis of antecedents and consequences of decision-making. Psychophysiology, 41, 531–540.

    Article  PubMed  Google Scholar 

  • Crowell, S. E., Beauchaine, T. P., Gatzke-Kopp, L., Sylvers, P., Mead, H., & Chipman-Chacon, J. (2006). Autonomic correlates of attention-deficit/hyperactivity and oppositional defiant disorder in preschool children. Journal of Abnormal Psychology, 115, 174–178.

    Article  PubMed  Google Scholar 

  • Crowne, D. P., & Marlowe, D. (1960). A new scale of social desirability independent of psychopathology. Journal of Consulting Psychology, 24, 349–354.

    Article  PubMed  Google Scholar 

  • Cruz, M. H., & Larsen, R. J. (1995). Personality correlates of individual differences in electrodermal lability. Social Behavior and Personality, 23, 93–104.

    Article  Google Scholar 

  • Curt, A., Weinhardt, C., & Dietz, V. (1996). Significance of sympathetic skin response in the assessment of autonomic failure in patients with spinal cord injury. Journal of the Autonomic Nervous System, 61, 175–180.

    Article  PubMed  Google Scholar 

  • Cuthbert, B. N., Bradley, M. M., & Lang, P. J. (1996). Probing picture perception: Activation and emotion. Psychophysiology, 33, 103–111.

    Article  PubMed  Google Scholar 

  • Damasio, A. R. (1994). Descartes Error: Emotion, reason and the human brain. New York: Avon.

    Google Scholar 

  • Damasio, A. R., Tranel, D., & Damasio, H. (1990). Individuals with sociopathic behavior caused by frontal damage fail to respond autonomically to social stimuli. Behavioural Brain Research, 41, 81–94.

    Article  PubMed  Google Scholar 

  • Damasio, A. R., Tranel, D., & Damasio, H. (1991). Somatic markers and the guidance of behavior: Theory and preliminary testing. In H. S. Levin, H. M. Eisenberg, & A. L. Benton (Eds.), Frontal lobe function and dysfunction (pp. 217–229). New York: Oxford University Press.

    Google Scholar 

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

    Google Scholar 

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

    Google Scholar 

  • Davidson, R. J. (1993). Cerebral asymmetry and emotion: Conceptual and methodological conundrums. Cognition and Emotion, 7, 115–138.

    Article  Google Scholar 

  • Davidson, R. A., Fedio, P., Smith, B. S., Aureille, E., & Martin, A. (1992). Lateralized mediation of arousal and habituation: Differential bilateral electrodermal activity in unilateral temporal lobectomy patients. Neuropsychologia, 30, 1053–1063.

    Article  PubMed  Google Scholar 

  • Davidson, R. A., & Smith, B. D. (1991). Caffeine and novelty: Effects on electrodermal activity and performance. Physiology and Behavior, 49, 1169–1175.

    Article  PubMed  Google Scholar 

  • Davies, D. R. (1983). Attention, arousal and effort. In A. Gale & J. A. Edwards (Eds.), Physiological correlates of human behaviour (Attention and performance, Vol. 2, pp. 9–34). London: Academic.

    Google Scholar 

  • Davis, M. (2000). The role of the amygdala in conditioned and unconditioned fear and anxiety. In J. P. Aggleton (Ed.), The amygdala: A functional analysis (2nd ed., pp. 213–287). Oxford: Oxford University Press.

    Google Scholar 

  • Davis, J. F., Malmo, R. B., & Shagass, C. (1954). Electromyographic reaction to strong auditory stimulation in psychiatric patients. Canadian Journal of Psychology, 8, 177–186.

    Article  PubMed  Google Scholar 

  • Dawson, M. E. (1980). Physiological detection of deception: Measurement of responses to questions and answers during countermeasure maneuvers. Psychophysiology, 17, 8–17.

    Article  PubMed  Google Scholar 

  • Dawson, M. E. (1990). Psychophysiology at the interface of clinical science, cognitive science, and neuroscience. Psychophysiology, 27, 243–255.

    Article  PubMed  Google Scholar 

  • Dawson, M. E., Catania, J. J., Schell, A. M., & Grings, W. W. (1979). Autonomic classical conditioning as a function of awareness of stimulus contingencies. Biological Psychology, 9, 23–40.

    Article  PubMed  Google Scholar 

  • Dawson, M. E., Filion, D. L., & Schell, A. M. (1989). Is elicitation of the autonomic orienting response associated with allocation of processing resources? Psychophysiology, 26, 560–572.

    Article  PubMed  Google Scholar 

  • Dawson, M. E., & Furedy, J. J. (1976). The role of awareness in human differential autonomic classical conditioning: The necessary gate hypothesis. Psychophysiology, 13, 50–53.

    Article  PubMed  Google Scholar 

  • Dawson, M. E., & Nuechterlein, K. H. (1984). Psychophysiological dysfunctions in the developmental course of schizophrenic disorders. Schizophrenia Bulletin, 10, 204–232.

    PubMed  Google Scholar 

  • Dawson, M. E., Nuechterlein, K. H., & Adams, R. M. (1989). Schizophrenic disorders. In G. Turpin (Ed.), Handbook of clinical psychophysiology (pp. 393–418). New York: Wiley.

    Google Scholar 

  • Dawson, M. E., Nuechterlein, K. H., Schell, A. M., Gitlin, M., & Ventura, J. (1994). Autonomic abnormalities in schizophrenia. Archives of General Psychology, 51, 813–824.

    Google Scholar 

  • Dawson, M. E., & Schell, A. M. (1982). Electrodermal responses to attended and nonattended significant stimuli during dichotic listening. Journal of Experimental Psychology. Human Perception and Performance, 8, 315–324.

    Article  PubMed  Google Scholar 

  • Dawson, M. E., & Schell, A. M. (1985). Information processing and human autonomic classical conditioning. In P. K. Ackles, J. R. Jennings, & M. G. H. Coles (Eds.), Advances in psychophysiology (Vol. 1, pp. 89–165). Greenwich, CT: Jai.

    Google Scholar 

  • Dawson, M. E., & Schell, A. M. (1987). Human autonomic and skeletal classical conditioning: The role of conscious cognitive factors. In G. Davey (Ed.), Cognitive processes and Pavlovian conditioning in humans (pp. 27–55). New York: Wiley.

    Google Scholar 

  • Dawson, M. E., & Schell, A. M. (2002). What does electrodermal activity tell us about prognosis in the schizophrenia spectrum? Schizophrenia Research, 54, 87–93.

    Article  PubMed  Google Scholar 

  • Dawson, M. E., Schell, A. M., Beers, J. R., & Kelly, A. (1982). Allocation of cognitive processing capacity during human autonomic classical conditioning. Journal of Experimental Psychology: General, 111, 272–295.

    Article  Google Scholar 

  • Dawson, M. E., Schell, A. M., & Catania, J. J. (1977). Autonomic correlates of depression and clinical improvement following electroconvulsive shock therapy. Psychophysiology, 14, 569–578.

    Article  PubMed  Google Scholar 

  • Dawson, M. E., Schell, A. M., & Filion, D. L. (1990). The electrodermal system. In I. T. Cacioppo & L. G. Tassinary (Eds.), Principles of psychophysiology (pp. 295–324). Cambridge: Cambridge University Press.

    Google Scholar 

  • De Bonis, M., & Freixa i Baqué, E. (1980). Stress, verbal cognitive activity and bilateral electrodermal responses. Neuropsychobiology, 6, 249–259.

    Article  PubMed  Google Scholar 

  • De Pascalis, V., Valerio, E., Santoro, M., & Cacace, I. (2007). Neuroticism-anxiety, impulsive-sensation seeking and autonomic responses to somatosensory stimuli. International Journal of Psychophysiology, 63, 16–24.

    Article  PubMed  Google Scholar 

  • De Seze, J., Stojkovic, T., Gauvrit, J.-Y., Devos, S., Ayachi, M., Cassim, F., et al. (2001). Autonomic dysfunction in multiple sclerosis: Cervical spinal cord atrophy correlates. Journal of Neurology, 248, 297–303.

    Article  PubMed  Google Scholar 

  • Dean, R. S. (1981). Lateral preference patterns as a discriminator of learning difficulties. Journal of Consulting and Clinical Psychology, 49, 227–235.

    Article  PubMed  Google Scholar 

  • DeLong, M. R., Georgopoulos, A. P., & Crutcher, M. D. (1983). Cortico-basal ganglia relations and coding of motor performance. Experimental Brain Research, Supplement, 7, 30–40.

    Google Scholar 

  • DeLong, M. R., & Wichmann, T. (2007). Circuits and circuit disorders of the basal ganglia. Archives of Neurology, 64, 20–24.

    Article  PubMed  Google Scholar 

  • Dembroski, T. M., MacDougall, J. M., & Shields, J. L. (1977). Physiologic reactions to social challenge in persons evidencing the type A coronary-prone behavior pattern. Journal of Human Stress, 3, 2–10.

    Article  PubMed  Google Scholar 

  • Dembroski, T. M., MacDougall, J. M., Shields, J. L., Petitto, J., & Lushene, R. (1978). Components of the type A coronary-prone behavior pattern and cardiovascular responses to psychomotor performance challenge. Journal of Behavioral Medicine, 1, 159–176.

    Article  PubMed  Google Scholar 

  • Dembroski, T. M., Weiss, S. M., Shields, J. L., Haynes, S. G., & Feinleib, M. (Eds.). (1978). Coronary prone behavior. New York: Springer.

    Google Scholar 

  • Denburg, N. L., Recknor, E. C., Bechara, A., & Tranel, D. (2006). Psychophysiological anticipation of positive outcomes promotes advantageous decision-making in normal older persons. International Journal of Psychophysiology, 61, 19–25.

    Article  PubMed  Google Scholar 

  • Dengerink, H. A., & Taylor, S. P. (1971). Multiple responses with differential properties in delayed galvanic skin response conditioning: A review. Psychophysiology, 8, 348–360.

    Article  PubMed  Google Scholar 

  • Denislic, M., & Meh, D. (1996). Sympathetic skin response in Parkinsonian patients. Electromyography and Clinical Neurophysiology, 36, 231–235.

    PubMed  Google Scholar 

  • Depue, R. A., & Fowles, D. C. (1973). Electrodermal activity as an index of arousal in schizophrenics. Psychological Bulletin, 79, 233–238.

    Article  PubMed  Google Scholar 

  • Derogatis, L. R., Klerman, G. L., & Lipman, R. S. (1972). Anxiety states and depressive neuroses. The Journal of Nervous and Mental Disease, 155, 392–403.

    Article  PubMed  Google Scholar 

  • Dettmers, C., van Ahlen, H., Faust, H., Fatepour, D., & Tackmann, W. (1994). Evaluation of erectile dysfunction with the sympathetic skin response in comparison to bulbocavernosus reflex and somatosensory evoked potentials of the pudendal nerve. Electromyography and Clinical Neurophysiology, 34, 437–444.

    PubMed  Google Scholar 

  • DiCara, L. V., & Miller, N. E. (1968). Instrumental learning of vasomotor responses by rats: Learning to respond differentially in the two ears. Science, 159, 1485–1486.

    Article  PubMed  Google Scholar 

  • Dimitriev, L., Belyakova, L., Bondarenko, T., & Nikolaev, G. (1968). Investigation of the orienting reaction and the defense reaction of schizophrenia in different stages of their illness. Zhurnal Nevropatologie Psikhiatrii, 68, 713–719.

    Google Scholar 

  • Dimond, S. J., & Beaumont, J. G. (1974). Experimental studies of hemisphere function in the human brain. In S. J. Dimond & J. G. Beaumont (Eds.), hemisphere function in the human brain (pp. 48–88). New York: Wiley.

    Google Scholar 

  • Dimond, S. J., Farrington, L., & Johnson, P. (1976). Differing emotional response from right and left hemispheres. Nature, 261, 690–692.

    Article  PubMed  Google Scholar 

  • Dindo, L., & Fowles, D. C. (2008). The skin conductance orienting response to semantic stimuli: Significance can be independent of arousal. Psychophysiology, 45, 111–118.

    PubMed  Google Scholar 

  • Doerr, H. O., Follette, W., Scribner, B. H., & Eisdorfer, C. (1980). Electrodermal response dysfunction in patients on maintenance renal dialysis. Psychophysiology, 17, 83–86.

    Article  PubMed  Google Scholar 

  • Dolu, N., Süer, C., Özesmi, C., Kelestimur, F., & Esel, E. (1997). Electrodermal activity in nonmedicated hyperthyroid patients having no depressive symptoms. Biological Psychiatry, 42, 1024–1029.

    Article  PubMed  Google Scholar 

  • Donat, D. C., & McCullough, J. P. (1983). Psychophysiological discriminants of depression at rest and in response to stress. Journal of Clinical Psychology, 39, 315–320.

    Article  PubMed  Google Scholar 

  • Douglas, R. J., & Pribram, K. H. (1966). Learning and limbic lesions. Neuropsychologia, 4, 197–220.

    Article  Google Scholar 

  • Dratcu, L., & Bond, A. (1998). Panic patients in the non-panic state: Physiological and cognitive dysfunction. European Psychiatry, 13, 18–25.

    Article  PubMed  Google Scholar 

  • Duffy, E. (1951). The concept of energy mobilization. Psychological Review, 58, 30–40.

    Article  PubMed  Google Scholar 

  • Duffy, E. (1972). Activation. In N. S. Greenfield & R. A. Sternbach (Eds.), Handbook of psychophysiology (pp. 577–622). New York: Holt, Rinehart, & Winston.

    Google Scholar 

  • Dunn, B. D., Dalgleish, T., & Lawrence, A. D. (2006). The somatic marker hypothesis: A critical evaluation. Neuroscience and Biobehavioral Reviews, 30, 239–271.

    Article  PubMed  Google Scholar 

  • Dykman, R. A., Reese, W. G., Galbrecht, C. R., & Thomasson, P. J. (1959). Psychophysiological reactions to novel stimuli: Measurement, adaptation, and relationship of psychological and physiological variables in the normal human. Annals of the New York Academy of Sciences, 79, 45–107.

    Article  Google Scholar 

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

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

    Google Scholar 

  • Edwards, J. A., & Siddle, D. A. T. (1976). Dishabituation of the electrodermal orienting response following decay of sensitization. Biological Psychology, 4, 19–28.

    Article  PubMed  Google Scholar 

  • Egyed, B., Eory, A., Veres, T., & Manninger, J. (1980). Measurement of electrical resistance after nerve injuries of the hand. The Hand, 12, 275–281.

    Article  PubMed  Google Scholar 

  • Eisfeld, W., Schaefer, F., Boucsein, W., & Stolz, C. (2005). Tracking intersensory properties of cosmetic products via psycho-physiological assessment. International Federation Societies of Cosmetic Chemists (IFSCC), 8, 25–30.

    Google Scholar 

  • Eisfeld, W., Wachter, R., Schaefer, F., & Boucsein, W. (2007). Objective emotional assessment of perceivable wellness effects. Cosmetics & Toiletries, 122, 63–72.

    Google Scholar 

  • Ekman, P., Friesen, W. V., & Ellsworth, P. C. (1972). Emotion in the human face: Guidelines for research and an integration of findings. New York: Plenum.

    Google Scholar 

  • Ekman, P., Friesen, W. V., & Simons, R. C. (1985). Is the startle reaction an emotion? Journal of Personality and Social Psychology, 49, 1416–1426.

    Article  PubMed  Google Scholar 

  • Ellingson, R. J. (1954). The incidence of EEG abnormality among patients with mental disorders of apparently nonorganic origin: A critical review. The American Journal of Psychology, 8, 263–275.

    Google Scholar 

  • Ellis, H. D., & de Pauw, K. W. (1994). The cognitive neuropsychiatric origins of the Capgras delusion. In A. S. David & J. C. Cutting (Eds.), The neuropsychology of schizophrenia (pp. 317–335). Hove: Psychology.

    Google Scholar 

  • Ellis, H. D., & Lewis, M. B. (2001). Capgras delusion: A window on face recognition. Trends in Cognitive Sciences, 5, 149–156.

    Article  PubMed  Google Scholar 

  • Ellis, H. D., Lewis, M. B., Moselhy, H. F., & Young, A. W. (2000). Automatic without autonomic responses to familiar faces: Differential components of covert face recognition in a case of capgras delusion. Cognitive Neuropsychiatry, 5, 255–269.

    Article  Google Scholar 

  • Ellis, H. D., & Young, A. W. (1990). Accounting for delusional misidentifications. The British Journal of Psychiatry, 157, 239–248.

    Article  PubMed  Google Scholar 

  • Ellis, H. D., Young, A. W., Quayle, A. H., & de Pauw, K. W. (1997). Reduced autonomic responses to faces in capgras delusion. Proceedings of the Royal Society of London, 264, 1085–1092.

    Article  Google Scholar 

  • Engel, B. T. (1972). Response specifity. In N. S. Greenfield & R. A. Sternbach (Eds.), Handbook of psychophysiology (pp. 571–576). New York: Holt, Rinehart, & Winston.

    Google Scholar 

  • Epstein, S. (1972). The nature of anxiety with emphasis upon its relationship to expectancy. In C. Spielberger (Ed.), Anxiety: Current trends in theory and research (Vol. 2, pp. 291–337). New York: Academic.

    Google Scholar 

  • Epstein, S., Boudreau, L., & Kling, S. (1975). Magnitude of the heart rate and electrodermal response as a function of stimulus input, motor output, and their interaction. Psychophysiology, 12, 15–24.

    Article  PubMed  Google Scholar 

  • Epstein, S., & Coleman, M. (1970). Drive theories of schizophrenia. Psychosomatic Medicine, 32, 113–140.

    PubMed  Google Scholar 

  • Erdmann, G., & Baumann, S. (1996). Sind psychophysiologische Veränderungen im Paradigma “Öffentliches Sprechen” Ausdruck emotionaler Belastung? Zeitschrift für Experimentelle Psychologie, 63, 224–255.

    Google Scholar 

  • Erdmann, G., Janke, W., & Bisping, R. (1984). Wirkungen und Vergleich der Wirkungen von vier experimentellen Belastungssituationen. Zeitschrift für Experimentelle und Angewandte Psychologie, 31, 521–543.

    Google Scholar 

  • Erdmann, G., Janke, W., Köchers, S., & Terschlüsen, B. (1984). Comparison of the emotional effects of a beta-adrenergic blocking agent and a tranquilizer under different situational conditions. I. Anxiety-arousing situations. Neuropsychobiology, 12, 143–151.

    Google Scholar 

  • Erlenmeyer-Kimling, L. (1975). A prospective study of children at risk for schizophrenia: Methodological considerations and some preliminary findings. In R. D. Wirt, G. Winokur, & M. Roff (Eds.), Life history research in psychopathology (Vol. 4, pp. 23–46). Minneapolis: University of Minnesota Press.

    Google Scholar 

  • Erlenmeyer-Kimling, L., Cornblatt, B., & Fleiss, J. (1979). High-risk research in schizophrenia. Psychiatric Annals, 9, 79–99.

    Google Scholar 

  • Erlenmeyer-Kimling, L., Marcuse, Y., Cornblatt, B., Friedman, D., Rainer, J. D., & Rutschmann, J. (1984). The New York high-risk project. In N. F. Watt, E. J. Anthony, L. C. Wynne, & J. E. Rolf (Eds.), Children at risk for schizophrenia: A longitudinal perspective (pp. 169–189). London: Cambridge University Press.

    Google Scholar 

  • Esen, F., Celebi, G., Ertekin, C., & Colakoglu, Z. (1997). Electrodermal activity in patients with Parkinson’s disease. Clinical Autonomic Research, 7, 35–40.

    Article  PubMed  Google Scholar 

  • Euteneuer, F., Schaefer, F., Stuermer, R., Boucsein, W., Timmermann, L., Barbe, M. T., et al. (2009). Dissociation of decision-making under ambiguity and decision-making under risk in patients with Parkinson’s disease: A neuropsychological and psychophysiological study. Neuropsychologia, 47, 2882–2890.

    Article  PubMed  Google Scholar 

  • Eysenck, H. J. (1957). Drugs and personality: I. Theory and methodology. Journal of Mental Science, 103, 119–131.

    PubMed  Google Scholar 

  • Eysenck, H. J. (1967). The biological basis of personality. Springfield: Thomas.

    Google Scholar 

  • Eysenck, M. W. (1982). Attention and arousal. Berlin: Springer.

    Google Scholar 

  • Eysenck, H. J. (1983). Psychophysiology and personality: Extraversion, neuroticism and psychoticism. In A. Gale & J. A. Edwards (Eds.), Physiological correlates of human behaviour (Individual differences and psychopathology, Vol. 3, pp. 13–30). London: Academic.

    Google Scholar 

  • Eysenck, H. J. (1994). Personality: Biological foundations. In P. A. Vernon (Ed.), The neuropsychology of individual differences (pp. 151–207). New York: Academic.

    Google Scholar 

  • Eysenck, H. J., & Eysenck, M. W. (1985). Personality and individual differences. New York: Plenum.

    Book  Google Scholar 

  • Eysenck, S., & Zuckerman, M. (1978). The relationship between sensation-seeking and Eysenck’s dimensions of personality. British Journal of Psychology, 69, 483–487.

    Article  Google Scholar 

  • Faber, S. (1983). Zur Auswertemethodik und Interpretation von Hautleitfähigkeitsmessungen bei arbeitswissenschaftlicher Beanspruchungsermittlung. Zeitschrift für Arbeitswissenschaft., 37(9 NF), 85–91.

    Google Scholar 

  • Fahrenberg, J. (1987). Concepts of activation and arousal in the theory of emotionality (neuroticism): A multivariate conceptualization. In J. Strelau & H. J. Eysenck (Eds.), Personality dimensions and arousal (pp. 99–120). New York: Plenum.

    Google Scholar 

  • Fahrenberg, J. (1988). Psychophysiological processes. In J. R. Nesselroade & R. B. Cattell (Eds.), Handbook of multivariate experimental psychology (pp. 867–914). New York: Plenum.

    Chapter  Google Scholar 

  • Fahrenberg, J., Walschburger, P., Foerster, F., Myrtek, M., & Müller, W. (1983). An evaluation of trait, state, and reaction aspects of activation processes. Psychophysiology, 20, 188–195.

    Article  PubMed  Google Scholar 

  • Falkowski, J., & Steptoe, A. (1983). Biofeedback-assisted relaxation in the control of reactions to a challenging task and anxiety-provoking film. Behaviour Research and Therapy, 21, 161–167.

    Article  PubMed  Google Scholar 

  • Farhoumand, N., Harrison, J., Pare, C. M. B., Turner, P., & Wynn, S. (1979). The effect of high dose oxprenolol on stress-induced physical and psychophysiological variables. Psychopharmacology, 64, 365–369.

    Article  PubMed  Google Scholar 

  • Feij, J. A. (1984). The psychophysiological and neurochemical bases of sensation seeking. In H. Bonarius, G. van Heck, & N. Smid (Eds.), Personality psychology in Europe (pp. 317–326). Lisse: Swets & Zeitlinger.

    Google Scholar 

  • Ficková, E. (1983). Dynamics of psychophysiological activation in shift-work operators. Studia Psychologica, 25, 105–113.

    Google Scholar 

  • Filion, D. L., Dawson, M. E., & Schell, A. M. (1998). The psychological significance of human startle eyeblink modification: A review. Biological Psychiatry, 47, 1–43.

    Article  Google Scholar 

  • Filion, D. L., Dawson, M. E., Schell, A. M., & Hazlett, E. A. (1991). The relationship between skin conductance orienting and the allocation of processing resources. Psychophysiology, 28, 410–424.

    Article  PubMed  Google Scholar 

  • Firth, H. (1973). Habituation during sleep. Psychophysiology, 10, 43–51.

    Article  PubMed  Google Scholar 

  • Fishbein, D., Eldreth, D., Matochik, J., Isenberg, I., Hyde, C., & London, E. (2005). Cognitive performance and autonomic reactivity in abstinent drug abusers and nonusers. Experimental and Clinical Psychopharmacology, 13, 25–40.

    Article  PubMed  Google Scholar 

  • Flor, H., Knost, B., & Birbaumer, N. (1997). Processing of pain- and body-related verbal material in chronic pain patients: Central and peripheral correlates. Pain, 73, 413–421.

    Article  PubMed  Google Scholar 

  • Flor-Henry, P. (1993). Electrodermal amplitude asymmetry and orienting response-non-response in psychopathology. In J.-C. Roy, W. Boucsein, C. D. Fowles, & J. H. Gruzelier (Eds.), Progress in Electrodermal Research (pp. 289–296). New York: Plenum.

    Chapter  Google Scholar 

  • Flykt, A., Esteves, F., & Öhman, A. (2007). Skin conductance response to masked conditioned stimuli: Phylogenetic/ontogenetic factors versus direction of threat? Biological Psychology, 74, 328–336.

    Article  PubMed  Google Scholar 

  • Folkins, C. H. (1970). Temporal factors and the cognitive mediators of stress reaction. Journal of Personality and Social Psychology, 14, 173–184.

    Article  PubMed  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 

  • Foulds, G. A., & Bedford, A. (1976). The relationship between anxiety-depression and the neuroses. The British Journal of Psychiatry, 128, 166–168.

    Article  PubMed  Google Scholar 

  • Fowler, C. J. (1993). Electrophysiologic evaluation of sexual dysfunction. In P. A. Low (Ed.), Clinical Autonomic Disorders (pp. 279–285). Boston: Little Brown.

    Google Scholar 

  • Fowler, R. L., & Kimmel, H. D. (1962). Operant conditioning of the GSR. Journal of Experimental Psychology, 63, 563–567.

    Article  PubMed  Google Scholar 

  • Fowles, D. C. (1980). The three arousal model: Implications of Gray’s two-factor learning theory for heart rate, electrodermal activity, and psychopathy. Psychophysiology, 17, 87–104.

    Article  PubMed  Google Scholar 

  • Fowles, D. C. (1986b). The psychophysiology of anxiety and hedonic affect: Motivational specificity. In B. F. Shaw, T. M. Segal, & T. M. Vallis (Eds.), Anxiety disorders (pp. 51–66). New York: Plenum.

    Google Scholar 

  • Fowles, D. C. (1988). Psychophysiology and psychopathology: A motivational approach. Psychophysiology, 25, 373–391.

    Article  PubMed  Google Scholar 

  • Fowles, D. C. (2000). Electrodermal hyporeactivity and antisocial behaviour: Does anxiety mediate the relationship? Journal of Affective Disorders, 61, 177–189.

    Article  PubMed  Google Scholar 

  • Fowles, D. C., Fisher, A. E., & Tranel, D. T. (1982). The heart beats to reward: The effect of monetary incentive on heart rate. Psychophysiology, 19, 506–513.

    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., Roberts, R., & Nagel, K. E. (1977). The influence of introversion/extraversion on the skin conductance response to stress and stimulus intensity. Journal of Research in Personality, 11, 129–146.

    Article  Google Scholar 

  • Franz, M., Olbrich, R., Croissant, B., Kirsch, P., Schmitz, N., & Schneider, C. (1999). Gefühl ohne Sprache oder Sprache ohne Gefühl? Weitere Hinweise auf die Validität der Entkopplungshypothese der Alexithymie. Nervenarzt, 70, 216–224.

    Article  PubMed  Google Scholar 

  • Fredrikson, M. (1981). Orienting and defensive reactions to phobic and conditioned fear stimuli in phobics and normals. Psychophysiology, 18, 456–465.

    Article  PubMed  Google Scholar 

  • Fredrikson, M., Annas, P., Georgiades, A., Hursti, T., & Tersman, Z. (1993). Internal consistency and temporal stability of classically conditioned skin conductance responses. Biological Psychology, 35, 153–163.

    Article  PubMed  Google Scholar 

  • Fredrikson, M., Dimberg, U., & Frisk-Holmberg, M. (1980). Arterial blood pressure and electrodermal activity in hypertensive and normotensive subjects during inner- and outer- directed attention. Acta Medica Scandinavica, 646, 73–76.

    Google Scholar 

  • Fredrikson, M., Dimberg, U., Frisk-Holmberg, M., & Ström, G. (1982). Haemodynamic and electrodermal correlates of psychogenic stimuli in hypertensive and normotensive subjects. Biological Psychology, 15, 63–73.

    Article  PubMed  Google Scholar 

  • Fredrikson, M., & Engel, B. T. (1985). Cardiovascular and electrodermal adjustments during a vigilance task in patients with borderline and established hypertension. Journal of Psychosomatic Research, 29, 235–246.

    Article  PubMed  Google Scholar 

  • Frith, C. D., & Allen, H. A. (1983). The skin conductance orienting response as an index of attention. Biological Psychology, 17, 27–39.

    Article  PubMed  Google Scholar 

  • Frith, C. D., Stevens, M., Johnstone, E. C., & Crow, T. J. (1979). Skin conductance responsivity during acute episodes of schizophrenia as a predictor of symptomatic improvement. Psychological Medicine, 9, 101–106.

    Article  PubMed  Google Scholar 

  • Frith, C. D., Stevens, M., Johnstone, E. C., & Crow, T. J. (1982). Skin conductance habituation during acute episodes of schizophrenia: Qualitative differences from anxious and depressed patients. Psychological Medicine, 12, 575–583.

    Article  PubMed  Google Scholar 

  • Frith, C. D., Stevens, M., Johnstone, E. C., & Owens, D. G. C. (1984). The effects of chronic treatment with amitriptyline and diazepam on electrodermal activity in neurotic outpatients. Physiological Psychology, 12, 247–252.

    Google Scholar 

  • Fuentes, I., Merita, M. G., Miquel, M., & Rojo, J. (1993). Relationships between electrodermal activity and symtomatology in schizophrenia. Psychopathology, 26, 47–52.

    Article  PubMed  Google Scholar 

  • Fung, M. T., Raine, A., Loeber, R., Lynam, D. R., Steinhauer, S. R., Venables, P. H., et al. (2005). Reduced electrodermal activity in psychopathy-prone adolescents. Journal of Abnormal Psychology, 114, 187–196.

    Article  PubMed  Google Scholar 

  • Furedy, J. J. (1967). Classical appetitive conditioning of the GSR with cool air as UCS, and the roles of UCS onset and offset as reinforcers of the CR. Journal of Experimental Psychology, 75, 73–80.

    Article  PubMed  Google Scholar 

  • Furedy, J. J. (1968). Human orienting reaction as a function of electrodermal-versus plethysmographic response modes and single versus alternating stimulus series. Journal of Experimental Psychology, 77, 70–78.

    Article  PubMed  Google Scholar 

  • Furedy, J. J. (1970). Test of the preparatory adaptive response interpretation of aversive classical autonomic conditioning. Journal of Experimental Psychology, 84, 301–307.

    Article  Google Scholar 

  • Furedy, J. J. (1972). Electrodermal recovery time as a supra sensitive autonomic index of anticipated intensity of threatened shock. Psychophysiology, 9, 281–282.

    Article  PubMed  Google Scholar 

  • Furedy, J. J. (1975). An integrative progress report on informational control in humans: Some laboratory findings and methodological claims. Australian Journal of Psychology, 27, 61–83.

    Article  Google Scholar 

  • Furedy, J. J. (1986). Lie detection as psychophysiological differentiation: Some fine lines. In M. G. H. Coles, E. Donchin, & S. W. Porges (Eds.), Psychophysiology: Systems, processes, and applications (pp. 683–701). Amsterdam: Elsevier.

    Google Scholar 

  • Furedy, J. J. (1987). Evaluating polygraphy from a psychophysiological perspective: A specific-effects analysis. The Pavlovian Journal of Biological Science, 22, 145–152.

    PubMed  Google Scholar 

  • Furedy, J. J., Arabian, J. M., Thiels, E., & George, L. (1982). Direct and continuous measurement of relational learning in human Pavlovian conditioning. The Pavlovian Journal of Biological Science, 17, 69–79.

    PubMed  Google Scholar 

  • Furedy, J. J., & Ben-Shakhar, G. (1991). The roles of deception, intention to deceive, and motivation to avoid detection in the psychophysiological detection of guilty knowledge. Psychophysiology, 28, 163–171.

    Article  PubMed  Google Scholar 

  • Furedy, J. J., Damke, B., & Boucsein, W. (2000). Revisiting the learning-without-awareness question in human Pavlovian autonomic conditioning: Focus on extinction in a dichotic listening paradigm. Integrative Physiological and Behavioral Science, 35, 17–34.

    Article  PubMed  Google Scholar 

  • Furedy, J. J., Davis, C., & Gurevich, M. (1988). Differentiation of deception as a psychological process: A psychophysiological approach. Psychophysiology, 25, 683–688.

    Article  PubMed  Google Scholar 

  • Furedy, J. J., Gigliotti, F., & Ben-Shakhar, G. (1994). Electrodermal differentiation of deception: The effect of choice vs. no choice of deceptive items. International Journal of Psychophysiology, 18, 13–22.

    Article  PubMed  Google Scholar 

  • Furedy, J. J., & Heslegrave, R. J. (1988). Validity of the lie detector: A psychphysiology perspective. Criminal Justice and Behavior, 15, 219–246.

    Article  Google Scholar 

  • Furedy, J. J., & Klajner, F. (1972). Unconfounded autonomic indexes of the aversiveness of signaled and unsignaled shocks. Journal of Experimental Psychology, 92, 313–318.

    Article  PubMed  Google Scholar 

  • Furedy, J. J., & Klajner, F. (1974). On evaluating autonomic and verbal indices of negative preception. Psychophysiology, 11, 121–124.

    Article  PubMed  Google Scholar 

  • Furedy, J. J., Posner, R. T., & Vincent, A. (1991). Electrodermal differentiation of deception: Perceived accuracy and perceived memorial content manipulations. International Journal of Psychophysiology, 11, 91–97.

    Article  PubMed  Google Scholar 

  • Furedy, J. J., & Poulos, C. X. (1977). Short-interval classical SCR conditioning and the stimulus-sequence-change-elicited OR: The case of the empirical red herring. Psychophysiology, 14, 351–359.

    Article  PubMed  Google Scholar 

  • Furedy, J. J., Poulos, C. X., & Schiffman, K. (1975). Contingency theory and classical autonomic excitatory and inhibitory conditioning: Some problems of assessment and interpretation. Psychophysiology, 12, 98–105.

    Article  PubMed  Google Scholar 

  • Furedy, J. J., & Riley, D. M. (1987). Human Pavlovian autonomic conditioning and the cognitive paradigm. In G. Davey (Ed.), Cognitive processes and Pavlovian conditioning in humans (pp. 1–25). New York: Wiley.

    Google Scholar 

  • Furedy, J. J., & Schiffman, K. (1971). Test of the propriety of the traditional discriminative control procedure in Pavlovian electrodermal and plethysmographic conditioning. Journal of Experimental Psychology, 91, 161–164.

    Article  PubMed  Google Scholar 

  • Furedy, J. J., & Schiffman, K. (1973). Concurrent measurement of autonomic and cognitive processes in a test of the traditional discriminative control procedure for Pavlovian electrodermal conditioning. Journal of Experimental Psychology, 100, 210–217.

    Article  PubMed  Google Scholar 

  • Furedy, J. J., & Schiffmann, K. (1974). Interrelationships between human classical differential electrodermal conditioning, orienting reaction, responsivity, and awareness of stimulus contingencies. Psychophysiology, 11, 58–67.

    Article  PubMed  Google Scholar 

  • Furedy, J. J., & Scull, J. (1971). Orienting-reaction theory and an increase in the human GSR following stimulus change which is unpredictable but not contrary to prediction. Journal of Experimental Psychology, 88, 292–294.

    Article  PubMed  Google Scholar 

  • Gaebelein, J., Taylor, S. P., & Borden, R. (1974). Effects of an external cue on psychophysiological reactions to a noxious event. Psychophysiology, 11, 315–320.

    Article  PubMed  Google Scholar 

  • Gainotti, G. (1979). The relationship between emotions and cerebral dominance: A review of clinical and experimental evidence. In J. H. Gruzelier & P. Flor-Henry (Eds.), Hemisphere asymmetries of function in psychopathology (pp. 21–34). Amsterdam: Elsevier.

    Google Scholar 

  • Gale, A., & Edwards, J. A. (Eds.). (1983). Physiological correlates of human behaviour (Three volumes). London: Academic.

    Google Scholar 

  • Gale, A., & Edwards, J. A. (1986). Individual differences. In M. G. H. Coles, E. Donchin, & W. Porges (Eds.), Psychophysiology: Systems, processes, and applications (pp. 431–507). Amsterdam: Elsevier.

    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 

  • Gatchel, R. J., McKinney, M. E., & Koebernick, L. F. (1977). Learned helplessness, depression, and physiological responding. Psychophysiology, 14, 25–31.

    Article  PubMed  Google Scholar 

  • Gati, I., & Ben-Shakhar, G. (1990). Novelty and significance in orientation and habituation: A feature-matching approach. Journal of Experimental Psychology. General, 119, 251–263.

    Article  PubMed  Google Scholar 

  • Geer, J. H., & Davison, G. C. (1970). Reduction of stress in humans through nonveridical perceived control of aversive stimulation. Journal of Personality and Social Psychology, 16, 731–738.

    Article  PubMed  Google Scholar 

  • Geer, J. H., & Maisel, E. (1972). Evaluating the effects of the prediction-control confound. Journal of Personality and Social Psychology, 23, 314–319.

    Article  PubMed  Google Scholar 

  • Gellhorn, E. (1964). Motion and emotion: The role of proprioception in the physiology and pathology of the emotions. Psychological Review, 71, 457–472.

    Article  PubMed  Google Scholar 

  • Giedke, H., & Bolz, J. (1980). Pre- and postimperative negative variation (CNV and PINV) under different conditions of controllability in depressed patients and healthy controls. In H. H. Kornhuber & L. Deecke (Eds.), Motivation, motor and sensory processes of the brain. Electrical potentials, behavior and clinical use (pp. 579–582). Amsterdam: Elsevier.

    Google Scholar 

  • Giedke, H., & Coenen, T. (1986). Die medikamentöse Behandlung von Angstzuständen. In W. Janke & P. Netter (Eds.), Angst und Psychopharmaka (pp. 207–234). Stuttgart: Kohlhammer.

    Google Scholar 

  • Goldberg, L. R. (1992). The development of markers for the Big-Five factor structure. Psychological Assessment, 4, 26–42.

    Article  Google Scholar 

  • Goldstein, I. B., & Shapiro, D. (1988). Cardiovascular responses to mental arithmetic and handgrip during different conditions of postural change. Psychophysiology, 25, 127–136.

    Article  PubMed  Google Scholar 

  • Gomez, R., & McLaren, S. (1997). The effects of reward and punishment on response disinhibition, moods, heart rate and skin conductance level during instrumental learning. Personality and Individual Differences, 23, 305–316.

    Article  Google Scholar 

  • Gormezano, I., & Kehoe, E. J. (1975). Classical conditioning: Some methodological-conceptual issues. In W. K. Estes (Ed.), Handbook of learning and cognitive processes (Conditioning and behavior theory, Vol. 2, pp. 143–179). Hillsdale, N.J.: Lawrence Erlbaum.

    Google Scholar 

  • Gougerot, M. L. (1947). Recherches sur l’impedance cutanée en courant alternatif de basse frequence au cours de differentes dermatoses. Annales et Bulletin de Dermatologie, 8, 101–111.

    Google Scholar 

  • Gough, H. (1969). Manual for the California psychological inventory. Palo Alto: Consulting Psychologists Press.

    Google Scholar 

  • Graham, F. K. (1973). Habituation and dishabituation of responses innervated by the autonomic nervous system. In H. V. S. Peeke & M. J. Herz (Eds.), Habituation (Behavioral studies, Vol. 1, pp. 163–218). New York: Academic.

    Google Scholar 

  • Graham, F. K. (1979). Distinguishing among orienting, defense, and startle reflexes. In H. D. Kimmel, E. H. van Olst, & J. F. Orlebeke (Eds.), The orienting reflex in humans (pp. 137–167). Hillsdale, NJ: Erlbaum.

    Google Scholar 

  • Gray, J. A. (1970). The psychophysiological basis of introversion-extraversion. Behaviour Research and Therapy, 8, 249–266.

    Article  PubMed  Google Scholar 

  • Gray, J. A. (1973). Causal theories of personality and how to test them. In J. R. Royce (Ed.), Multivariate analysis and psychological theory (pp. 409–463). New York: Academic.

    Google Scholar 

  • Gray, J. A. (1981). A critique of Eysenck’s theory of personality. In H. J. Eysenck (Ed.), A model for personality (pp. 246–276). New York: Springer.

    Chapter  Google Scholar 

  • Gray, J. A. (1982). The neuropsychology of anxiety: An inquiry into the functions of the septo-hippocampal system. Oxford: Clarendon.

    Google Scholar 

  • Gray, J. A. (1987). A conceptual nervous system for avoidance behaviour. In J. A. Gray (Ed.), The psychology of fear and stress (pp. 241–331). Cambridge: University Press.

    Google Scholar 

  • Gray, J. A., & McNaughton, N. (2000). The neuropsychology of anxiety (2nd ed.). New York: Oxford University Press.

    Google Scholar 

  • Gray, J. A., & Smith, P. T. (1969). An arousal-decision model for partial reinforcement and discrimination learning. In R. Gilbert & N. S. Sutherland (Eds.), Animal discrimination learning (pp. 243–272). New York: Academic.

    Google Scholar 

  • Greenwald, M. K., Cook, E. W., & Lang, P. J. (1989). Affective judgement and psychophysiological response: Dimensional covariation in the evaluation of pictorial stimuli. Journal of Psychophysiology, 3, 51–64.

    Google Scholar 

  • Grillon, C. (2002). Associative learning deficits increase symptoms of anxiety in humans. Biological Psychiatry, 51, 851–858.

    Article  PubMed  Google Scholar 

  • Grillon, C., & Davis, M. (1997). Fear-potentiated startle conditioning in humans: Explicit and contextual cue conditioning following paired vs. unpaired training. Psychophysiology, 34, 451–458.

    Article  PubMed  Google Scholar 

  • Grings, W. W. (1960). Preparatory set variables related to classical conditioning of autonomic responses. Psychological Review, 67, 243–252.

    Article  PubMed  Google Scholar 

  • Grings, W. W. (1969). Anticipatory and preparatory electrodermal behavior in paired stimulation situations. Psychopysiology, 5, 597–611.

    Article  Google Scholar 

  • Grings, W. W., & Dawson, M. E. (1973). Complex variables in conditioning. In W. Prokasy & D. C. Raskin (Eds.), Electrodermal activity in psychological research (pp. 203–254). New York: Academic.

    Google Scholar 

  • Grings, W. W., & Dawson, M. E. (1978). Emotions and bodily responses: A psychophysiological approach. New York: Academic.

    Google Scholar 

  • Grings, W. W., Givens, M. C., & Carey, C. A. (1979). Contingency contrast effects in discrimination conditioning. Journal of Experimental Psychology. General, 108, 281–295.

    Article  PubMed  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 

  • Groeppel-Klein, A. (2005). Arousal and consumer in-store behavior. Brain Research Bulletin, 67, 428–437.

    Article  PubMed  Google Scholar 

  • Groeppel-Klein, A., Domke, A., & Bartmann, B. (2006). Pretty woman or Erin Brokovich? Unconscious and conscious reactions to commercials and movies shaped by fairy tale archetypes – Results from two experimental studies. Advances in Consumer Research, 33, 163–174.

    Article  Google Scholar 

  • Groeppel-Klein, A., Germelmann, C., & Woratschek, H. (2007). Arousal and consumer decision-making. Proceedings of the 36th European marketing academy conference. Reykjavik, Iceland.

    Google Scholar 

  • Gross, J. J. (1998). Antecedent- and response-focused emotion regulation: Divergent consequences for experience, expression, and physiology. Journal of Personality and Social Psychology, 74, 224–237.

    Article  PubMed  Google Scholar 

  • Gross, J. J., Carstensen, L. L., Pasupathi, M., Tsai, J., Skorpen, C. G., & Hsu, A. Y. C. (1997). Emotion and aging: Experience, expression and control. Psychology and Aging, 12, 590–599.

    Article  PubMed  Google Scholar 

  • Gross, J. J., & Levenson, R. W. (1993). Emotional suppression: Physiology, self-report, and expressive behavior. Journal of Personality and Social Psychology, 64, 970–986.

    Article  PubMed  Google Scholar 

  • Gross, J. J., & Levenson, R. W. (1997). Hiding feelings: The acute effects of inhibiting negative and positive emotion. Journal of Abnormal Psychology, 106, 95–103.

    Article  PubMed  Google Scholar 

  • Gross, C. G., & Sergent, J. (1992). Face recognition. Current Opinion in Neurobiology, 2, 156–161.

    Article  PubMed  Google Scholar 

  • Groves, P. M., & Thompson, R. F. (1970). Habituation: A dual-process theory. Psychological Review, 77, 419–450.

    Article  PubMed  Google Scholar 

  • Gruzelier, J. H. (1973). Bilateral asymmetry of skin conductance orienting activity and levels in schizophrenics. Biological Psychology, 1, 21–41.

    Article  PubMed  Google Scholar 

  • Gruzelier, J. H. (1976). Clinical attributes of schizophrenic skin conductance responders and nonresponders. Psychological Medicine, 6, 245–249.

    Article  PubMed  Google Scholar 

  • Gruzelier, J. H. (1979). Lateral asymmetries in electrodermal activity and psychosis. In J. H. Gruzelier & P. Flor-Henry (Eds.), Hemisphere asymmetries of function in psychopathology (pp. 701–713). Amsterdam: Elsevier.

    Google Scholar 

  • Gruzelier, J. H. (1983). Disparate syndromes in psychosis delineated by direction of electrodermal response lateral asymmetry. In P. Flor-Henry & J. Gruzelier (Eds.), Laterality and psychopathology (pp. 525–538). Amsterdam: Elsevier.

    Google Scholar 

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

    Google Scholar 

  • Gruzelier, J. H., & Connolly, J. F. (1979). Differential drug action on electrodermal orienting responses as distinct from nonspecific responses and electrodermal levels. In H. D. Kimmel, E. H. van Olst, & J. F. Orlebeke (Eds.), The orienting reflex in humans (pp. 701–713). Hillsdale, NJ: Erlbaum.

    Google Scholar 

  • Gruzelier, J. H., Eves, F., & Connolly, J. (1981a). Reciprocal hemispheric influences on response habituation in the electrodermal system. Physiological Psychology, 9, 313–317.

    Google Scholar 

  • Gruzelier, J. H., Eves, F., Connolly, J. F., & Hirsch, S. R. (1981c). Orienting, habituation, sensitization, and dishabituation in the electrodermal system of consecutive, drug-free admissions for schizophrenia. Biological Psychology, 12, 187–209.

    Article  PubMed  Google Scholar 

  • Gruzelier, J. H., & Hammond, N. V. (1976). Schizophrenia: A dominant hemisphere temporal-limbic disorder? Research Communications in Psychology, Psychiatry and Behavior, 1, 32–72.

    Google Scholar 

  • Gruzelier, J. H., & Hammond, N. V. (1977). The effect of chlorpromazine upon bilateral asymmetries in bioelectrical skin reactivity of schizophrenics. Studia Psychologica, 19, 40–51.

    Google Scholar 

  • Gruzelier, J. H., & Hammond, N. V. (1978). The effect of chlorpromazine upon psychophysiological, endocrine and information processing measures in schizophrenia. Journal of Psychiatric Research, 14, 167–182.

    Article  PubMed  Google Scholar 

  • Gruzelier, J. H., & Manchanda, R. (1982). The syndrome of schizophrenia: Relations between electrodermal response, lateral asymmetries and clinical ratings. The British Journal of Psychiatry, 141, 488–495.

    Article  PubMed  Google Scholar 

  • Gruzelier, J. H., Nixon, P. G. F., Liddiard, D., Pugh, S., & Baxter, R. (1986). Retarded habituation and lateral asymmetries in electrodermal activity in cardiovascular disorders. International Journal of Psychophysiology, 3, 219–226.

    Article  PubMed  Google Scholar 

  • Gruzelier, J., Sergeant, J., & Eves, F. (1988). The use of bilateral skin conductance measurement in elucidating stimulus versus response processing influences on the orienting reaction. International Journal of Psychophysiology, 6, 195–205.

    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 

  • Gruzelier, J. H., & Venables, P. H. (1973). Skin conductance responses to tones with and without attentional significance in schizophrenic and nonschizophrenic psychiatric patients. Neuropsychologia, 11, 221–230.

    Article  PubMed  Google Scholar 

  • Gruzelier, J. H., & Venables, P. H. (1974). Bimodality and lateral asymmetry of skin conductance orienting activity in schizophrenics: Replication and evidence of lateral asymmetry in patients with depression and disorders of personality. Biological Psychiatry, 8, 55–73.

    PubMed  Google Scholar 

  • Gudjonsson, G. H. (1981). Self-reported emotional disturbance and its relation to electrodermal reactivity; defensiveness and trait anxiety. Personality and Individual Differences, 2, 47–52.

    Article  Google Scholar 

  • Gudjonsson, G. H. (1986). The validity of polygraph techniques in lie detection. In D. Papakostopoulos, S. Butler, & I. Martin (Eds.), Clinical and experimental neuropsychophysiology (pp. 448–465). Dover: Croom Helm.

    Google Scholar 

  • Guidotti, A., Baraldi, M., Schwartz, J. P., & Costa, E. (1979). Molecular mechanisms regulating the interaction between benzodiazepines and GABA receptors in the central nervous system. Pharmacology Biochemistry and Behavior, 10, 803–807.

    Article  Google Scholar 

  • Guinjoan, S. M., Bernabó, J. L., & Cardinali, D. P. (1995). Cardiovascular test of autonomic function and sympathetic skin responses in patients with major depression. Journal of Neurology, Neuosurgery, and Psychiatry, 58, 299–302.

    Article  Google Scholar 

  • Gutrecht, J. A., Suarez, G. A., & Denny, B. E. (1993). Sympathetic skin response in multiple sclerosis. Journal of the Neurological Sciences, 118, 88–91.

    Article  PubMed  Google Scholar 

  • Haapaniemi, T. H., Korpelainen, J. T., Tolonen, U., Suominen, K., Sotaniemi, K. A., & Myllylä, V. V. (2000). Suppressed sympathetic skin response in Parkinson disease. Clinical Autonomic Research, 10, 337–342.

    Article  PubMed  Google Scholar 

  • Haarmann, A., Boucsein, W., & Schaefer, F. (2009). Combining electrodermal responses and cardiovascular measures for probing adaptive automation during simulated flight. Applied Ergonomics, 40, 1026–1040.

    Article  PubMed  Google Scholar 

  • Hagemann, T., Levenson, R. W., & Gross, J. J. (2006). Expressive suppression during an acoustic startle. Psychophysiology, 43, 104–112.

    Article  PubMed  Google Scholar 

  • Haider, M. (1969). Elektrophysiologische Indikatoren der Aktiviertheit. In W. Schönpflug (Ed.), Methoden der Aktivierungsforschung (pp. 125–156). Bern: Huber.

    Google Scholar 

  • Haider, M. (1970). Neuropsychology of attention, expectation, and vigilance. In D. I. Mostofski (Ed.), Attention: Contemporary theory and analysis (pp. 419–432). New York: Appleton-Century-Crofts.

    Google Scholar 

  • Hamann, S., Monarch, E. S., & Goldstein, F. C. (2002). Impaired fear conditioning in Alzheimer’s disease. Neuropsychologia, 40, 1187–1195.

    Article  PubMed  Google Scholar 

  • Hanisch, L. J., Palmer, S. C., Donahue, A., & Coyne, J. C. (2007). Validation of sternal skin conductance for detection of hot flashes in prostate cancer survivors. Psychophysiology, 44, 189–193.

    Article  PubMed  Google Scholar 

  • Hare, R. D. (1975). Psychopathy. In P. H. Venables & M. J. Christie (Eds.), Research in psychophysiology (pp. 325–348). London: Wiley.

    Google Scholar 

  • Hare, R. D. (1978a). Psychopathy and electrodermal responses to nonsignal stimulation. Biological Psychology, 6, 237–246.

    Article  PubMed  Google Scholar 

  • Hare, R. D. (1978b). Electrodermal and cardiovascular correlates of psychopathy. In R. D. Hare & D. Schalling (Eds.), Psychopathic behaviour: Approaches to research (pp. 107–143). New York: Wiley.

    Google Scholar 

  • Harris, M. D. (1943). Habituatory response decrement in the intact organism. Psychological Bulletin, 40, 385–422.

    Article  Google Scholar 

  • Hart, J. D. (1974). Physiological responses of anxious and normal subjects to simple signal and non-signal auditory stimuli. Psychophysiology, 11, 443–451.

    Article  PubMed  Google Scholar 

  • Hastrup, J. L. (1979). Effects of electrodermal lability and introversion on vigilance decrement. Psychophysiology, 16, 302–310.

    Article  PubMed  Google Scholar 

  • Hastrup, J. L., & Katkin, E. S. (1976). Electrodermal lability: An attempt to measure its psychological correlates. Psychophysiology, 13, 296–301.

    Article  PubMed  Google Scholar 

  • Hazlett, E. A., Dawson, M. E., Buchsbaum, M. S., & Nuechterlein, K. H. (1993). Reduced regional brain metabolism assessed by positron emission tomography in electrodermal nonresponder schizophrenics: A pilot study. Journal of Abnormal Psychology, 102, 39–46.

    Article  PubMed  Google Scholar 

  • Healey, J. A., & Picard, R. W. (2005). Detecting stress during real-world driving tasks using physiological sensors. IEEE Transactions on Intelligent Transportation Systems, 156–166.

    Google Scholar 

  • Hécaen, H., & Sauguet, J. (1971). Cerebral dominance in left handed subjects. Cortex, 7, 19–48.

    PubMed  Google Scholar 

  • Heimann, H. (1969). Typologische und statistische Erfassung depressiver Syndrome. In H. Hippius & H. Selbach (Eds.), Das depressive Syndrom: Internationales Symposium, Berlin 1968 (pp. 279–290). München: Urban & Schwarzenberg.

    Google Scholar 

  • Heimann, H. (1978). Changes of psychophysiological reactivity in affective disorders. Archiv für Psychiatrie und Nervenkrankheiten, 225, 223–231.

    Article  PubMed  Google Scholar 

  • Heimann, H. (1979). Auf dem Wege zu einer einheitlichen psychophysiologischen Theorie depressiver Syndrome. Praxis der Psychotherapie und Psychosomatik, 24, 281–297.

    Google Scholar 

  • Heimann, H. (1980). Psychophysiologische Aspekte in der Depressionsforschung. In H. Heimann & H. Giedke (Eds.), Neue Perspektiven in der Depressionsforschung (pp. 85–87). Bern: Huber.

    Google Scholar 

  • Heimer, L., Van Hoesen, G. W., Trimble, M., & Zahm, D. S. (2008). Anatomy of neuropsychiatry: The new anatomy of the basal forebrain and its implications for neuropsychiatric illness. London: Academic.

    Google Scholar 

  • Helander, M. (1974). Drivers’ physiological reactions and control operations as influenced by traffic events. Zeitschrift für Verkehrssicherheit, 20, 174–187.

    Google Scholar 

  • Helander, M. (1978). Applicability of drivers’ electrodermal response to the design of the traffic environment. Journal of Applied Psychology, 63, 481–488.

    Article  PubMed  Google Scholar 

  • Helmer, J. E., & Furedy, J. J. (1968). Operant conditioning of GSR amplitude. Journal of Experimental Psychology, 78, 463–467.

    Article  PubMed  Google Scholar 

  • Herpertz, S. C., Mueller, B., Qunaibi, M., Lichterfeld, C., Konrad, K., & Herpertz-Dahlmann, B. (2005). Response to emotional stimuli in boys with conduct disorder. The American Journal of Psychiatry, 162, 1100–1107.

    Article  PubMed  Google Scholar 

  • Hiroshige, Y., & Iwahara, S. (1978). Digital and cephalic vasomotor orienting responses to indifferent, signal, and verbal stimuli. Psychophysiology, 15, 226–232.

    Article  PubMed  Google Scholar 

  • Hirstein, W., & Ramachandran, V. S. (1997). Capgras syndrome: A novel probe for understanding the neural representation of the identity and familiarity of persons. Proceedings of the Royal Society of London, 264, 437–444.

    Article  Google Scholar 

  • Hodges, W. E. (1976). The psychophysiology of anxiety. In M. Zuckerman & C. D. Spielberger (Eds.), Emotions and anxiety: New concepts, methods, and applications (pp. 175–194). Hillsdale, NJ: Erlbaum.

    Google Scholar 

  • Hoefer, M., Allison, S. C., Schauer, G. G., Neuhaus, J. M., Hall, J., Dang, J. N., et al. (2008). Fear conditioning in Alzheimer’s disease. Brain, 131, 1646–1657.

    Article  PubMed  Google Scholar 

  • Holloway, F. A., & Parsons, O. A. (1969). Unilateral brain damage and bilateral skin conductance levels in humans. Psychophysiology, 6, 138–148.

    Article  PubMed  Google Scholar 

  • Holmes, D. S., Frost, D. O., Bennett, D. H., Nielsen, D. H., & Lutz, D. J. (1981). Effectiveness of skin resistance biofeedback for controlling arousal in non-stressful and stressful situations: Two experiments. Journal of Psychosomatic Research, 25, 205–211.

    Article  PubMed  Google Scholar 

  • Holmes, D. S., McGilley, B. M., & Houston, B. K. (1984). Task-related arousal of type A and type B persons: Level of challenge and response specifity. Journal of Personality and Social Psychology, 46, 1322–1327.

    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 

  • Honts, C. R. (2004). The psychophysiological detection of deception. In P. Granhag & L. Strömwall (Eds.), Detection of deception in forensic contexts (pp. 103–123). London: Cambridge University Press.

    Chapter  Google Scholar 

  • Honts, C. R., & Amato, S. (2007). Automation of a screening polygraph test increases accuracy. Psychology, Crime and Law, 13, 187–199.

    Article  Google Scholar 

  • Honts, C. R., Raskin, D. C., & Kircher, J. C. (1987). Effects of physical countermeasures and their electromyographic detection during polygraph tests for deception. Journal of Psychophysiology, 1, 241–247.

    Google Scholar 

  • Hori, T. (1982). Electrodermal and electro-oculographic activity in a hypnagogic state. Psychophysiology, 19, 668–672.

    Article  PubMed  Google Scholar 

  • Hot, P., Leconte, P., & Sequeira, H. (2005). Diurnal autonomic variations and emotional reactivity. Biological Psychology, 69, 261–270.

    Article  PubMed  Google Scholar 

  • Hot, P., Naveteur, J., Leconte, P., & Sequeira, H. (1999). Diurnal variations of tonic electrodermal activity. International Journal of Psychophysiology, 33, 223–230.

    Article  PubMed  Google Scholar 

  • Houston, B. K. (1983). Psychophysiological responsivity and the type A behavior pattern. Journal of Research in Personality, 17, 22–39.

    Article  Google Scholar 

  • Hu, S., McChesney, K. A., Player, K. A., Bahl, A. M., Buchanan, J. B., & Scozzafava, J. E. (1999). Systematic investigation of physiological correlates of motion sickness induced by viewing an optokinetic rotating drum. Aviation, Space, and Environmental Medicine, 70, 759–765.

    PubMed  Google Scholar 

  • Huber, D., Henrich, G., & Gündel, H. (2005). Psychophysiological response patterns of migraine patients in two habituation tests. Headache, 45, 1375–1387.

    Article  PubMed  Google Scholar 

  • Huck, S. W., & McLean, R. A. (1975). Using repeated measures ANOVA to analyse the data from pretest-posttest design: A potentially confusing task. Psychological Bulletin, 82, 511–518.

    Article  Google Scholar 

  • Hugdahl, K. (1984). Hemispheric asymmetry and bilateral electrodermal recordings: A review of the evidence. Psychophysiology, 21, 371–393.

    Article  PubMed  Google Scholar 

  • Hugdahl, K. (1988). Bilateral electrodermal asymmetry: Past hopes and future prospects. International Journal of Neuroscience, 39, 33–44.

    Article  PubMed  Google Scholar 

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

    Google Scholar 

  • Hugdahl, K. (1998). Cortical control of human classical conditioning: Autonomic and positron emission tomography data. Psychophysiology, 35, 170–178.

    Article  PubMed  Google Scholar 

  • Hugdahl, K., Broman, J.-E., & Franzon, M. (1983). Effects of stimulus content and brain lateralization on the habituation of the electrodermal orienting reaction (OR). Biological Psychology, 17, 153–168.

    Article  PubMed  Google Scholar 

  • Hugdahl, K., & Öhman, A. (1980). Skin conductance conditioning to potentially phobic stimuli as a function of interstimulus interval and delay versus trace paradigm. Psychophysiology, 17, 348–355.

    Article  PubMed  Google Scholar 

  • Humphrey, G. (1933). The nature of learning. New York: Harcourt Brace.

    Google Scholar 

  • Iaboni, F., Douglas, V. I., & Ditto, B. (1997). Psychophysiological response of ADHD children to reward and extinction. Psychophysiology, 34, 116–123.

    Article  PubMed  Google Scholar 

  • Iacono, W. G. (1982). Bilateral electrodermal habituation-dishabituation and resting EEG in remitted schizophrenics. The Journal of Nervous and Mental Disease, 170, 91–101.

    Article  PubMed  Google Scholar 

  • Iacono, W. G. (1985). Psychophysiologic markers of psychophathology: A review. Canadian Psychology, 26, 96–112.

    Article  Google Scholar 

  • Iacono, W. G. (1991). Psychophysiological assessment of psychopathology. Psychological Assessment, 3, 309–320.

    Article  Google Scholar 

  • Iacono, W. G. (2007). Detection of deception. In J. T. Cacioppo, L. G. Tassinary, & G. Berntson (Eds.), Handbook of psychophysiology (pp. 688–703). New York: Cambridge University Press.

    Google Scholar 

  • Iacono, W. G., Boisvenu, G. A., & Fleming, J. A. (1984). Effects of diazepam and methylphenidate on the electrodermal detection of guilty knowledge. Journal of Applied Psychology, 69, 289–299.

    Article  PubMed  Google Scholar 

  • Iacono, W. G., Ficken, J. W., & Beiser, M. (1999). Electrodermal activation in first-episode psychotic patients in their first-degree relatives. Psychiatry Research, 88, 25–39.

    Article  PubMed  Google Scholar 

  • Iacono, W. G., & Lykken, D. T. (1979). The orienting response: Importance of instructions. Schizophrenia Bulletin, 5, 11–14.

    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 

  • Iacono, W. G., Lykken, D. T., Peloquin, L. J., Lumry, A. E., Valentine, R. H., & Tuason, V. B. (1983). Electrodermal activity in euthymic unipolar and bipolar affective disorders: A possible marker for depression. Archives of General Psychiatry, 40, 557–565.

    PubMed  Google Scholar 

  • Iacono, W. G., Roshi, D., & Lacoste, D. (1987). Electrodermal activity in patients with Huntington’s disease and their progeny. Psychophysiology, 24, 522–527.

    Article  PubMed  Google Scholar 

  • Iacono, W. G., & Tuason, V. B. (1983). Bilateral electrodermal asymmetry in euthymic patients with unipolar and bipolar affective disorders. Biological Psychiatry, 18, 303–315.

    PubMed  Google Scholar 

  • Izard, C. E. (1971). Face of emotion. New York: Appleton.

    Google Scholar 

  • Jackson, J. C. (1974). Amplitude and habituation of the orienting reflex as a function of stimulus intensity. Psychophysiology, 11, 647–658.

    Article  PubMed  Google Scholar 

  • Jamner, L. D., & Tursky, B. (1987). Syndrome-specific descriptor profiling: A psychophysiological and psychophysical approach. Health Psychology, 6, 417–430.

    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., & Stern, J. A. (1976). Electrodermal response configuration as a function of rated psychopathology in children. The Journal of Nervous and Mental Disease, 162, 184–194.

    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 

  • Janke, W., & Netter, P. (1986). Angstbeeinflussung durch Psychopharmaka: Methodische Ansätze und Grundprobleme. In W. Janke & P. Netter (Eds.), Angst und Psychopharmaka (pp. 43–71). Stuttgart: Kohlhammer.

    Google Scholar 

  • Janssen, S. A., Arntz, A., & Bouts, S. (1998). Anxiety and pain: Epinephrine-induced hyperalgesia and attentional influences. Pain, 76, 309–316.

    Article  PubMed  Google Scholar 

  • Jarrett, A. (Ed.). (1978). The physiology and pathophysiology of the skin (The sweat glands, skin permeation, lymphatics, and the nails, Vol. 5). New York: Academic.

    Google Scholar 

  • Jenkinson, P. M., Baker, S. R., Edelstyn, N. M. J., & Ellis, S. J. (2008). Does autonomic arousal distinguish good and bad decisions? Journal of Psychophysiology, 22, 141–149.

    Article  Google Scholar 

  • Johns, M. W., Cornell, B. A., & Masterton, J. P. (1969). Monitoring sleep of hospital patients by measurement of electrical resistance of skin. Journal of Applied Psychology, 27, 898–901.

    Google Scholar 

  • Johnsen, B. H., Thayer, J. F., & Hugdahl, K. (1995). Affective judgment of the Ekman faces: A dimensional approach. Journal of Psychophysiology, 9, 193–202.

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

    Article  PubMed  Google Scholar 

  • Johnson, L. C., & Lubin, A. (1967). The orienting reflex during waking and sleeping. Electroencephalography and Clinical Neurophysiology, 22, 11–21.

    Article  PubMed  Google Scholar 

  • Johnson, H. J., & Schwartz, G. E. (1967). Suppression of GSR activity through operant reinforcement. Journal of Experimental Psychology, 75, 307–312.

    Article  PubMed  Google Scholar 

  • Johnson, L. C., Townsend, R. E., & Wilson, M. R. (1975). Habituation during sleeping and waking. Psychophysiology, 12, 574–584.

    Article  PubMed  Google Scholar 

  • Johnstone, E. C., Bourne, R. C., Crow, T. J., Frith, C. D., Gamble, S., Lofthouse, R., et al. (1981). The relationships between clinical response, psychophysiological variables and plasma levels of amitriptyline and diazepam in neurotic outpatients. Psychopharmacology, 72, 233–240.

    Article  PubMed  Google Scholar 

  • Jones, J. E. (1962). Contiguity and reinforcement in relation to CS-UCS intervals in classical aversive conditioning. Psychological Review, 69, 176–186.

    Article  PubMed  Google Scholar 

  • Jörg, J., & Boucsein, W. (1998). Die sympathische Hautreaktion (SSR). Klinische Neurophysiologie, 29, 186–197.

    Article  Google Scholar 

  • Jörg, J., Jock, S., Boucsein, W., & Schäfer, F. (2004). Zur autonomen Dysregulation beim Freezing-Phänomen von Morbus-Parkinson-Patienten. Ein ambulatorisches Monitoring und Videorecording. Aktuelle Neurologie, 31, 338–346.

    Article  Google Scholar 

  • Jovanovic, U. J. (1971). Normal sleep in man. Stuttgart: Hippokrates.

    Google Scholar 

  • Jutai, J. W., & Hare, R. D. (1983). Psychopathy and selective attention during performance of a complex perceptual-motor task. Psychophysiology, 20, 146–151.

    Article  PubMed  Google Scholar 

  • Kahabka, G., Oppelt, W., Rohmert, W., & Müller, D. (1986). Geforderter Pilot – Gestreßter Fluggast. Die Beanspruchung von Pilot und Passagier beim Motorflug. Aerokurier, 3, 274–276.

    Google Scholar 

  • Kahneman, D. (1973). Attention and effort. Englewood Cliffs: Prentice-Hall.

    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 

  • Katkin, E. S. (1975). Electrodermal lability: A psychophysiological analysis of individual differences in response to stress. In C. D. Spielberger & I. G. Sarason (Eds.), Stress and anxiety (Vol. 2, pp. 141–176). New York: Wiley.

    Google Scholar 

  • Katkin, E. S., & Deitz, S. R. (1973). Systematic desensitization. In W. F. Prokasy & D. C. Raskin (Eds.), Electrodermal activity in psychological research (pp. 347–376). New York: Academic.

    Google Scholar 

  • Katkin, E. S., & McCubbin, R. J. (1969). Habituation of the orienting response as a function of individual differences in anxiety and autonomic lability. Journal of Abnormal Psychology, 74, 54–60.

    Article  PubMed  Google Scholar 

  • Katkin, E. S., & Murray, E. N. (1968). Instrumental conditioning of autonomically mediated behavior: Theoretical and methodological issues. Psychological Bulletin, 70, 52–68.

    Article  PubMed  Google Scholar 

  • Katsanis, J., & Iacono, W. G. (1992). Temporal lobe dysfunction and electrodermal nonresponding in schizophrenia. Biological Psychiatry, 31, 159–170.

    Article  PubMed  Google Scholar 

  • Katz, R. (1984). Unconfounded electrodermal measures in assessing the aversiveness of predictable and unpredictable shocks. Psychophysiology, 21, 452–458.

    Article  PubMed  Google Scholar 

  • Katz, R., & Wykes, T. (1985). The psychological difference between temporally predictable and unpredictable stressful events: Evidence for information control theories. Journal of Personality and Social Psychology, 48, 781–790.

    Article  Google Scholar 

  • Kayser, J. (1995). Hemisphärenunterschiede, Emotion und bilaterale elektrodermale Aktivität. Frankfurt am Main: Peter Lang.

    Google Scholar 

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

    Google Scholar 

  • Kendler, H. H. (1952). “What is learned?” – A theoretical blind alley. Psychological Review, 59, 269–277.

    Article  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 

  • Ketterer, M. W., & Smith, B. D. (1982). Lateralized cortical/cognitive processing and electrodermal activity: Effects of subject and stimulus characteristics. Psychophysiology, 19, 328–329.

    Google Scholar 

  • Kilpatrick, D. G. (1972). Differential responsiveness of two electrodermal indices to psychological stress and performance of a complex cognitive task. Psychophysiology, 9, 218–226.

    Article  PubMed  Google Scholar 

  • Kim, K. H., Bang, S. W., & Kim, S. R. (2004). Emotion recognition system using short-term monitoring of physiological signals. Medical & Biological Engineering & Computing, 42, 419–427.

    Article  Google Scholar 

  • Kim, D. K. K., Shin, M. Y., Kim, C. E., Cho, H. S., & Kim, Y. S. (1993). Electrodermal responsiveness, clinical variables, and brain imaging in male chronic schizophrenics. Biological Psychiatry, 33, 786–793.

    Article  PubMed  Google Scholar 

  • Kimble, G. A. (1961). Hilgard and Marquis’ conditioning and learning. New York: Appleton-Century-Crofts.

    Google Scholar 

  • Kimmel, H. D. (1966). Inhibition of the unconditioned response in classical conditioning. Psychological Review, 73, 232–240.

    Article  PubMed  Google Scholar 

  • Kimmel, H. D. (1967). Instrumental conditioning of autonomically mediated behavior. Psychological Bulletin, 67, 337–345.

    Article  PubMed  Google Scholar 

  • Kimmel, H. D. (1973). Instrumental conditioning. In W. F. Prokasy & D. C. Raskin (Eds.), Electrodermal activity in psychological research (pp. 255–282). New York: Academic.

    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 

  • Kimmel, H. D., van Olst, E. H., & Orlebeke, J. F. (Eds.). (1979). The orienting reflex in humans. Hillsdale, NJ: Erlbaum.

    Google Scholar 

  • Kimura, D. (1969). Spatial localization in left and right visual field. Canadian Journal of Psychology, 23, 445–458.

    Article  PubMed  Google Scholar 

  • Kimura, D. (1973). The asymmetry of the human brain. Scientific American, 228, 70–78.

    Article  PubMed  Google Scholar 

  • Kircher, J. C., & Raskin, D. C. (1988). Human versus computerized evaluations of polygraph data in a laboratory setting. Journal of Applied Psychology, 73, 291–302.

    Article  PubMed  Google Scholar 

  • Kirsch, P., & Boucsein, W. (1997). Classical conditioning and information processing: Different mechanism for prepared and unprepared stimuli? Integrative Physiological and Behavioral Science, 32, 247–256.

    Article  PubMed  Google Scholar 

  • Kirsch, P., Boucsein, W., & Baltissen, R. (1993). Electrodermal activity as an indicator of information processing in a nonaversive differential classical conditioning paradigm. Integrative Physiological and Behavioral Science, 28, 154–157.

    Article  PubMed  Google Scholar 

  • Kiss, G. (1979). Messung der elektrischen Impedanz zur Bestimmung von durch Laugen bedingten Hautschädigungen. Dermatologische Monatsschrift, 165, 526–530.

    Google Scholar 

  • Kiss, G., Horvath, I., & Hajdu, B. (1975). Elektrische Meßmethode und Gerät zum Nachweis bösartiger Wucherungen der Haut. Dermatologische Monatsschrift, 161, 374–378.

    Google Scholar 

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

    Google Scholar 

  • Kleck, R. E., Vaughan, R. C., Cartwright-Smith, J., Vaughan, K. B., Colby, C. Z., & Lanzetta, J. T. (1976). Effects of being observed on expressive, subjective, and physiological responses to painful stimuli. Journal of Personality and Social Psychology, 34, 1211–1218.

    Article  PubMed  Google Scholar 

  • Kleeberg, J., Bruggimann, L., Annoni, J.-M., van Melle, G., Bogousslavsky, J., & Schluep, M. (2004). Altered decision-making in multiple sclerosis: A sign of impaired emotional reactivity? Annals of Neurology, 56, 787–795.

    Article  PubMed  Google Scholar 

  • Klein, R. G., Abikoff, H., Klass, E., Ganeles, D., Seese, L. M., & Pollack, S. (1997). Clinical efficacy of methylphenidate in conduct disorder with and without attention deficit hyperactivity disorder. Archives of General Psychiatry, 54, 1073–1080.

    PubMed  Google Scholar 

  • Klein, D. F., & Rabkin, J. (Eds.). (1981). Anxiety: New research and changing concepts. New York: Raven.

    Google Scholar 

  • Kleitman, N. (1963). Sleep and wakefulness. Chicago: University of Chicago Press.

    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 

  • Knygazev, G. G., Slobodskaya, H. R., & Wilson, G. D. (2002). Psychophysiological correlates of behavioural inhibition and activation. Personality and Individual Differences, 33, 647–660.

    Article  Google Scholar 

  • Koehler, T., & Weber, D. (1992). Psychophysiological reactions of patients with atopic dermatitis. Journal of Psychosomatic Research, 36, 391–394.

    Article  PubMed  Google Scholar 

  • Koelega, H. S. (1990). Vigilance performance: A review of electrodermal predictors. Perceptual and Motor Skills, 70, 1011–1029.

    Google Scholar 

  • Koella, W. P. (1986). Psycho – und neuropharmakologische Wirkungen und Wirkungsmechanismen von Anxiolytika vom Benzodiazepin – und Beta-Rezeptorenblocker-Typ. In W. Janke & P. Netter (Eds.), Angst und Psychopharmaka (pp. 73–90). Stuttgart: Kohlhammer.

    Google Scholar 

  • Koepke, J. E., & Pribram, K. H. (1966). Habituation of GSR as a function of stimulus duration and spontaneous activity. Journal of Comparative and Physiological Psychology, 61, 442–448.

    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 

  • Koller, M., Zidek, H., & Haider, M. (1986). Induced psychophysiological stress reactions in patients suffering from myocardial infarction and peptic ulcer. Activitas Nervosa Superior, 28, 123–128.

    PubMed  Google Scholar 

  • Kopp, M. S. (1984). Electrodermal characteristics in psychosomatic patients groups. International Journal of Psychophysiology, 2, 73–85.

    Article  PubMed  Google Scholar 

  • Koriat, A., Averill, J. R., & Malmstrom, E. J. (1973). Individual differences in habituation: Some methodological and conceptual issues. Journal of Research in Personality, 7, 88–101.

    Article  Google Scholar 

  • Kornhuber, H. H., & Deecke, L. (1965). Hirnpotentialänderungen bei Willkürbewegungen und passiven Bewegungen des Menschen: Bereitschaftspotential und reafferente Potentiale. Pflügers Archiv für die gesamte Physiologie, 284, 1–17.

    Article  Google Scholar 

  • Korpelainen, J. T., Tolonen, U., Sotaniemi, K. A., & Myllylä, V. V. (1993). Suppressed sympathetic skin response in brain infarction. Stroke, 24, 1389–1392.

    Article  PubMed  Google Scholar 

  • Kotses, H., Rapaport, I., & Glaus, K. D. (1978). Operant conditioning of skin resistance tonic levels. Biofeedback and Self-Regulation, 3, 43–50.

    Article  PubMed  Google Scholar 

  • Koumans, A. J. R., Tursky, B., & Solomon, P. (1968). Electrodermal levels and fluctuations during normal sleep. Psychophysiology, 5, 300–306.

    Article  PubMed  Google Scholar 

  • Krantz, D. S., Glass, D. C., & Snyder, M. L. (1974). Helplessness, stress level, and the coronary-prone behavior pattern. Journal of Experimental Social Psychology, 10, 284–300.

    Article  Google Scholar 

  • Krapohl, D. J., Senter, S. M., & Stern, B. A. (2005). An exploration of methods for analysis of multiple-issue relevant/irrelevant screening data. Polygraph, 34, 47–61.

    Google Scholar 

  • Kreibig, S. D., Wilhelm, F. H., Roth, W. T., & Gross, J. J. (2007). Cardiovascular, electrodermal, and respiratory response patterns to fear- and sadness-inducing films. Psychophysiology, 44, 787–806.

    Article  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 

  • Kroeber-Riel, W. (1979). Activation research: Psychobiological approaches in consumer research. Journal of Consumer Research, 5, 240–250.

    Article  Google Scholar 

  • Kroeber-Riel, W., & Weinberg, P. (2003). Konsumentenverhalten. München: Vahlen.

    Google Scholar 

  • Krupski, A., Raskin, D. C., & Bakan, P. (1971). Physiological and personality correlates of commission errors in an auditory vigilance task. Psychophysiology, 8, 304–311.

    Article  PubMed  Google Scholar 

  • Kugelmass, S., Faber, N., Ingraham, L. J., Frenkel, E., Nathan, M., Mirsky, A. F., et al. (1995). Reanalysis of SCOR and anxiety measures in the Israeli high-risk study. Schizophrenia Bulletin, 21, 205–217.

    PubMed  Google Scholar 

  • Kugler, B. T., & Gruzelier, J. H. (1980). The influence of chlorpromazine and amylobarbitone on the recovery limb of the electrodermal response. Psychiatry Research, 2, 75–84.

    Article  PubMed  Google Scholar 

  • Kuhmann, W. (1989). Experimental investigation of stress-inducing properties of system response times. Ergonomics, 31, 271–280.

    Article  Google Scholar 

  • Kuhmann, W., Boucsein, W., Schaefer, F., & Alexander, J. (1987). Experimental investigation of psychophysiological stress-reactions induced by different system response times in human-computer interaction. Ergonomics, 30, 933–943.

    Article  PubMed  Google Scholar 

  • Kuhmann, W., Schaefer, F., & Boucsein, W. (1990). Effekte von Wartezeiten innerhalb einfacher Aufgaben: Eine Analogie zu Wartezeiten in der Mensch-Computer-Interaktion. Zeitschrift für Experimentelle und Angewandte Psychologie, 37, 242–265.

    PubMed  Google Scholar 

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

    Google Scholar 

  • Kushniruk, A., Rustenburg, J., & Ogilvie, R. (1985). Psychological correlates of electrodermal activity during REM sleep. Sleep, 8, 146–154.

    PubMed  Google Scholar 

  • LaBar, K. S., & Phelps, E. A. (1998). Arousal-mediated memory consolidation: Role of the medial temporal lobe in humans. Psychological Science, 9, 490–493.

    Article  Google Scholar 

  • LaBarbera, P. A., & Tucciarone, J. D. (1995). GSR reconsidered: A behavior-based approach to evaluating and improving the sales potency of advertising. Journal of Advertising Research, 35, 33–53.

    Google Scholar 

  • Lacey, J. I. (1967). Somatic response patterning and stress: Some revisions of activation theory. In M. H. Appley & R. Trumbull (Eds.), Psychological stress: Issues in research (pp. 14–37). New York: Appleton-Century-Crofts.

    Google Scholar 

  • Lacey, J. I., & Lacey, B. C. (1958). The relationship of resting autonomic activity to motor impulsivity. Research Publications – Association for Nervous and Mental Diseases, 36, 144–209.

    Google Scholar 

  • Lacey, J. I., & Lacey, B. C. (1970). Some autonomic-central nervous system interrelationships. In P. Black (Ed.), Physiological correlates of emotion (pp. 205–227). New York: Academic.

    Google Scholar 

  • Lacey, B. C., & Lacey, J. I. (1974). Studies of heart rate and other bodily processes in sensorimotor behavior. In P. A. Obrist, A. H. Black, J. Brener, & L. V. DiCara (Eds.), Cardiovascular psychophysiology (pp. 538–564). Chicago: Aldine-Atherton.

    Google Scholar 

  • Lacroix, J. M., & Comper, P. (1979). Lateralization in the electrodermal system as a function of cognitive/hemispheric manipulations. Psychophysiology, 16, 116–129.

    Article  PubMed  Google Scholar 

  • Lader, M. H. (1964). The effect of cyclobarbitone on the habituation of the psychogalvanic reflex. Brain, 87, 321–340.

    Article  PubMed  Google Scholar 

  • Lader, M. H. (1967). Palmar skin conductance measures in anxiety and phobic states. Journal of Psychosomatic Research, 11, 271–281.

    Article  PubMed  Google Scholar 

  • Lader, M. H. (1975). The psychophysiology of mental illness. London: Routledge.

    Google Scholar 

  • Lader, M. H. (1979). Anxiety reducing and sedation: A psychophysiological theory. British Journal of Clinical Pharmacology, 7, 91–105.

    Google Scholar 

  • Lader, M. H., & Mathews, A. M. (1970). Physiological changes during spontaneous panic attacks. Journal of Psychosomatic Research, 14, 377–382.

    Article  PubMed  Google Scholar 

  • Lader, M. H., & Petursson, H. (1983). Rational use of anxiolytic/sedative drugs. Drugs, 25, 514–528.

    Article  PubMed  Google Scholar 

  • Lader, M. H., & Wing, L. (1964). Habituation of the psycho-galvanic reflex in patients with anxiety states and in normal subjects. Journal of Neurology, Neurosurgery, and Psychiatry, 27, 210–218.

    Article  PubMed  Google Scholar 

  • Lader, M. H., & Wing, L. (1966). Physiological measures, sedative drugs, and morbid anxiety. London: Oxford University Press.

    Google Scholar 

  • Lader, M. H., & Wing, L. (1969). Physiological measures in agitated and retarded depressed patients. Journal of Psychiatric Research, 7, 89–100.

    Article  PubMed  Google Scholar 

  • Landis, C., & Hunt, W. A. (1939). The startle pattern. New York: Farrar & Rinehart.

    Google Scholar 

  • Lang, P. J. (1970). Stimulus control, response control and the desensitization of fear. In D. J. Lewis (Ed.), Learning approaches to therapeutic behavior change (pp. 148–173). Chicago: Aldine.

    Google Scholar 

  • Lang, P. J., Bradley, M. M., & Cuthbert, B. N. (1990). Emotion, attention, and the startle reflex. Psychological Review, 97, 377–395.

    Article  PubMed  Google Scholar 

  • Lang, P. J., Bradley, M. M., & Cuthbert, B. N. (1999). International Affective Picture System (IAPS): Digitized photographs, instruction manual and affective ratings. Technical Report A-6. Gainsville, FL: University of Florida.

    Google Scholar 

  • Langosch, W., Brodner, G., & Foerster, F. (1983). Psychophysiological testing of postinfarction patients: A study determining the cardiological importance of psychophysiological variables. In T. M. Dembroski, T. H. Schmidt, & G. Blümchen (Eds.), Biobehavioral bases of coronary heart disease (pp. 197–227). Basel: Karger.

    Google Scholar 

  • Lanzetta, J. T., Cartwright-Smith, J., & Kleck, R. E. (1976). Effects of nonverbal dissimulation on emotional experience and autonomic arousal. Journal of Personality and Social Psychology, 33, 354–370.

    Article  PubMed  Google Scholar 

  • Lapierre, Y. D. (1975). Clinical and physiological assessment of chlorazepate, diazepam and placebo in anxious neurotics. International Journal of Clinical Pharmacology, 11, 315–322.

    Google Scholar 

  • Lapierre, Y. D., & Butter, H. J. (1980). Agitated and retarded depression: A clinical psychophysiological evaluation. Neuropsychology, 6, 217–223.

    Google Scholar 

  • Latzman, R. D., Knutson, J. F., & Fowles, D. C. (2006). Schedule-induced electrodermal responding in children. Psychophysiology, 43, 623–632.

    Article  PubMed  Google Scholar 

  • Lawler, K. (1980). Cardiovascular and electrodermal response patterns in heart rate reactive individuals during psychological stress. Psychophysiology, 17, 464–470.

    Article  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 

  • Lawson, E. A. (1981). Skin conductance responses in Huntington’s chorea progeny. Psychophysiology, 18, 32–35.

    Article  PubMed  Google Scholar 

  • Lazarus, R. S. (1966). Psychological stress and the coping process. New York: McGraw-Hill.

    Google Scholar 

  • Lazarus, R. S., & Opton, E. M. (1966). The study of psychological stress: A summary of theoretical formulations and empirical findings. In C. D. Spielberger (Ed.), Anxiety and behavior (pp. 225–262). New York: Academic.

    Google Scholar 

  • Lazzaro, I., Gordon, E., Li, W., Lim, C. L., Plahn, M., Whitmont, S., et al. (1999). Simultaneous EEG and EDA measures in adolescent attention deficit hyperactivity disorder. International Journal of Psychophysiology, 34, 123–134.

    Article  PubMed  Google Scholar 

  • Ledowski, T., Preuss, J., Ford, A., Paech, M. J., McTernan, C., Kapila, R., et al. (2007). New parameters of skin conductance compared with bispectral index® monitoring to assess emergence from total intravenous anaesthesia. British Journal of Anaesthesia, 99, 547–551.

    Article  PubMed  Google Scholar 

  • Lenhart, R. E. (1985). Lowered skin conductance in a subsyndromal high-risk depressive sample: Response amplitudes versus tonic levels. Journal of Abnormal Psychology, 94, 649–652.

    Article  PubMed  Google Scholar 

  • Léonard, T., Pepinà, C., Bond, A., & Treasure, J. (1998). Assessment of test-meal induced autonomic arousal in anorexic, bulimic and control females. European Eating Disorders Review, 6, 188–200.

    Article  Google Scholar 

  • Leonard, J. P., Podoll, K., Weiler, H.-T., & Lange, H. W. (1984). Habituation der elektrodermalen Orientierungsreaktion in der Diagnostik und Früherkennung der Chorea Huntington. Zeitschrift für Experimentelle und Angewandte Psychologie, 31, 447–463.

    PubMed  Google Scholar 

  • Lester, B. K., Burch, N. R., & Dossett, R. C. (1967). Nocturnal EEG-GSR profiles: The influence of presleep states. Psychophysiology, 3, 238–248.

    Article  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 

  • Levenson, R. W. (1988). Emotion and the autonomic nervous system: A prospectus for research on autonomic specificity. In H. L. Wagner (Ed.), Social psychophysiology and emotion: Theory and clinical applications (pp. 17–42). New York: Wiley.

    Google Scholar 

  • Levenson, R. W., Carstensen, L. L., Friesen, W. V., & Ekman, P. (1991). Emotion, physiology, and expression in old age. Physiology and Aging, 6, 28–35.

    Google Scholar 

  • Levenson, R. W., Ekman, P., & Friesen, W. V. (1990). Voluntary facial action generates emotion specific autonomic nervous system activity. Psychophysiology, 27, 363–384.

    Article  PubMed  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 

  • Levinson, D. F., Edelberg, R., & Bridger, W. H. (1984). The orienting response in schizophrenia: Proposed resolution of a controversy. Biological Psychiatry, 19, 489–507.

    PubMed  Google Scholar 

  • Levis, D. J., & Smith, J. E. (1987). Getting individual differences in autonomic reactivity to work for instead of against you: Determining the dominant “psychological” stress channel on the basis of a “biological” stress test. Psychophysiology, 24, 346–352.

    Article  PubMed  Google Scholar 

  • Lim, C. L., Gordon, E., Harris, A., Bahramali, H., Li, W. M., Manor, B., et al. (1999). Electrodermal activity in schizophrenia: A quantitative study using a short interstimulus paradigm. Biological Psychiatry, 45, 127–135.

    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 

  • Linden, D., & Berlit, P. (1995). Sympathetic skin responses (SSRs) in monofocal brain lesions: Topographical aspects of central sympathetic pathways. Acta Neurologica Scandinavica, 91, 372–376.

    Article  PubMed  Google Scholar 

  • Lindholm, E., & Cheatham, C. M. (1983). Autonomic activity and workload during learning of a simulated aircraft carrier landing task. Aviation, Space, and Environmental Medicine, 54, 435–439.

    PubMed  Google Scholar 

  • Lindsley, D. B. (1951). Emotion. In S. S. Stevens (Ed.), Handbook of experimental psychology (pp. 473–516). New York: Wiley.

    Google Scholar 

  • Lindsley, D. B., Schreiner, L. H., Knowles, W. B., & Magoun, H. W. (1950). Behavioral and EEG changes following chronic brainstem lesions in the cat. Electroencephalography and Clinical Neurophysiology, 2, 483–498.

    Article  PubMed  Google Scholar 

  • Lipp, O. V., Siddle, D. A. T., & Dall, P. J. (1997). The effect of emotional and attentional processes on blink startle modulation and on electrodermal responses. Psychophysiology, 34, 340–347.

    Article  PubMed  Google Scholar 

  • Loeb, J., & Mednick, S. A. (1977). A prospective study of predictors of criminality: Electrodermal response patterns. In S. A. Mednick & K. O. Christiansen (Eds.), Biosocial bases of criminal behavior (pp. 245–254). New York: Gardener.

    Google Scholar 

  • Lorber, M. F. (2004). Psychophysiology of aggression, psychopathy, and conduct problems: A meta-analysis. Psychological Bulletin, 130, 531–552.

    Article  PubMed  Google Scholar 

  • Lovallo, W. R., & Pishkin, V. (1980). A psychophysiological comparison of Type A and B men exposed to failure and uncontrollable noise. Psychophysiology, 17, 29–36.

    Article  PubMed  Google Scholar 

  • Lovibond, P. F., Siddle, D. A. T., & Bond, N. W. (1993). Resistance to extinction of fear-relevant stimuli: Preparedness or selective sensitization? Journal of Experimental Psychology, 122, 449–461.

    Article  PubMed  Google Scholar 

  • Lykken, D. T. (1957). A study of anxiety in the sociopathic personality. Journal of Abnormal Psychology, 55, 6–10.

    Article  PubMed  Google Scholar 

  • Lykken, D. T. (1959b). The GSR in the detection of guilt. The Journal of Applied Psychology, 43, 385–388.

    Article  Google Scholar 

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

    Google Scholar 

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

    Article  Google Scholar 

  • Lykken, D. T. (1981). A tremor in the blood: Uses and abuses of the lie detector. New York: McGraw-Hill.

    Google Scholar 

  • Lykken, D. T. (1995). The antisocial personalities. Hillsdale, NJ: Lawrence Erlbaum.

    Google Scholar 

  • Lykken, D. T., Macindoe, I., & Tellegen, A. (1972). Preception: Autonomic response to shock as a function of predictability in time and locus. Psychophysiology, 9, 318–333.

    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., 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., & Tellegen, A. (1974). On the validity of the preception hypothesis. Psychophysiology, 11, 125–132.

    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 

  • Lynn, R. (1966). Attention, arousal and the orientation reaction. Oxford: Pergamon.

    Google Scholar 

  • Lyskov, E., Sandström, M., & Mild, K. H. (2001). Neurophysiological study of patients with perceived “electrical hypersensivity”. International Journal of Psychophysiology, 42, 233–241.

    Article  PubMed  Google Scholar 

  • Lyvers, M., & Miyata, Y. (1993). Effects of cigarette smoking on electrodermal orienting reflexes to stimulus change and stimulus significance. Psychophysiology, 30, 231–236.

    Article  PubMed  Google Scholar 

  • MacIntosh, B. J., McIlroy, W. E., Mraz, R., Staines, R., Black, S. E., & Graham, S. J. (2008). Electrodermal recording and fMRI to inform sensorimotor recovery in stroke patients. Neurorehabilitation and Neural Repair, 22, 728–736.

    Article  PubMed  Google Scholar 

  • Macleod, A. F., Smith, S. A., Cowell, T., Richardson, P. R., & Sonksen, P. H. (1991). Non-cardiac autonomic test in diabetes: Use of the galvanic skin response. Diabetic Medicine, 8 Symposium, S67-S70.

    Google Scholar 

  • Magaro, P. A. (1973). Skin conductance basal level and reactivity in schizophrenia as a function of chronicity, premorbid adjustment, diagnosis, and medication. Journal of Abnormal Psychology, 81, 270–281.

    Article  PubMed  Google Scholar 

  • Magliero, A., Gatchel, R. J., & Lojewski, D. (1981). Skin conductance responses to stimulus “energy” decreases following habituation. Psychophysiology, 18, 549–558.

    Article  PubMed  Google Scholar 

  • Maia, T. V., & McClelland, J. L. (2004). A reexamination of the evidence for the somatic marker hypothesis: What participants really know in the Iowa Gambling task. Proceedings of the National Academy of Sciences USA, 101, 16075–16080.

    Article  Google Scholar 

  • Malmo, R. B. (1957). Anxiety and behavioral arousal. Psychological Review, 64, 309–319.

    Article  Google Scholar 

  • Malmo, R. B. (1959). Activation: A neuropsychological dimension. Psychological Review, 66, 367–386.

    Article  PubMed  Google Scholar 

  • Malmo, R. B. (1962). Activation. In A. J. Bachrach (Ed.), Experimental foundations of clinical psychology (pp. 386–422). New York: Basic Books.

    Google Scholar 

  • Malmo, R. B., & Smith, A. A. (1951). Responsiveness in chronic schizophrenia. Journal of Personality, 18, 359–375.

    Article  Google Scholar 

  • Malten, K. E., & Thiele, F. A. J. (1973). Evaluation of skin damage. The British Journal of Dermatology, 89, 565–569.

    Article  PubMed  Google Scholar 

  • Maltzman, I. (1977). Orienting in classical conditioning and generalization of the galvanic skin response to words: An overview. Journal of Experimentel Psychology: General, 106, 111–119.

    Article  Google Scholar 

  • Maltzman, I. (1979a). Orienting reflexes and significance: A reply to O’Gorman. Psychophysiology, 16, 274–282.

    Article  PubMed  Google Scholar 

  • Maltzman, I. (1979b). Orienting reflexes and classical conditioning in humans. In H. D. Kimmel, E. H. van Olst, & J. F. Orlebeke (Eds.), The orienting reflex in humans (pp. 323–351). Hillsdale, NJ: Erlbaum.

    Google Scholar 

  • Maltzman, I., & Langdon, B. (1982). Novelty and significance as determiners of the GSR index of the orienting reflex. Physiological Psychology, 10, 229–234.

    Google Scholar 

  • Maltzman, I., Raskin, D. C., & Wolff, C. (1979). Latent inhibition of the GSR conditioned to words. Physiological Psychology, 7, 193–203.

    Google Scholar 

  • Mangan, G. L., & O’Gorman, J. G. (1969). Initial amplitude and rate of habituation of orienting reaction in relation to extraversion and neuroticism. Journal of Experimental Research in Personality, 3, 275–282.

    Google Scholar 

  • Mangina, C. A., & Beuzeron-Mangina, J. H. (1996). Direct electrical stimulation of specific human brain structures and bilateral electrodermal activity. International Journal of Psychophysiology, 22, 1–8.

    Article  PubMed  Google Scholar 

  • Maricq, H. R., & Edelberg, R. (1975). Electrodermal recovery rate in a schizophrenic population. Psychophysiology, 12, 630–633.

    Article  PubMed  Google Scholar 

  • Martin, R. B., & Dean, S. J. (1970). Instrumental modification of the GSR. Psychophysiology, 7, 178–185.

    Article  PubMed  Google Scholar 

  • Martin, A., & Rief, W. (2009). Wie wirksam ist Biofeedback? Eine therapeutische Methode. Bern: Hans Huber.

    Google Scholar 

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

    Article  PubMed  Google Scholar 

  • Massaro, D. W. (1975). Experimental psychology and information processing. Chicago: Rand McNally.

    Google Scholar 

  • Mayer, B., Merckelbach, H., de Jong, P. J., & Leeuw, I. (1999). Skin conductance responses of spider phobics to backwardly masked phobic cues. Journal of Psychophysiology, 13, 152–159.

    Article  Google Scholar 

  • Maze, J. R. (1983). The Meaning of Behaviour. London: Allen & Unwin.

    Google Scholar 

  • McCubbin, R. J., & Katkin, E. S. (1971). Magnitude of the orienting response as a function of extent and quality of stimulus change. Journal of Experimental Psychology, 88, 182–188.

    Article  PubMed  Google Scholar 

  • McDonald, D. G., & Carpenter, F. A. (1975). Habituation of the orienting response in sleep. Psychophysiology, 12, 618–623.

    Article  PubMed  Google Scholar 

  • McDonald, D. G., Shallenberger, H. D., Koresko, R. L., & Kinzy, B. G. (1976). Studies of spontaneous electrodermal responses in sleep. Psychophysiology, 13, 128–134.

    Article  PubMed  Google Scholar 

  • McDougall, A. J., & McLeod, J. G. (2003). Autonomic nervous system function in multiple sclerosis. Journal of the Neurological Sciences, 215, 79–85.

    Article  PubMed  Google Scholar 

  • McFarland, R. A. (1981). Physiological psychology: The biology of human behavior. Palo Alto: Mayfield.

    Google Scholar 

  • McGrath, J. E. (1982). Methodological problems in research on stress. In H. W. Krohne & L. Laux (Eds.), Achievement, stress, and anxiety (pp. 19–48). Washington: Hemisphere.

    Google Scholar 

  • McGuinness, D. (1973). Cardiovascular responses during habituation and mental activity in anxious men and women. Biological Psychology, 1, 115–124.

    Article  PubMed  Google Scholar 

  • McGuinness, D., & Pribram, K. (1980). The neuropsychology of attention: Emotional and motivational controls. In M. C. Wittrock (Ed.), The brain and psychology (pp. 95–139). New York: Academic.

    Google Scholar 

  • McKeever, W. F., & Gill, K. M. (1972). Interhemispheric transfer time for visual stimulus information varies as a function of the retinal locus of stimulation. Psychonomic Science, 26, 308–310.

    Google Scholar 

  • Mednick, S. A. (1967). The children of schizophrenics: Serious difficulties in current research methodologies which suggest the use of the “high-risk group” method. In J. Romano (Ed.), Origins of schizophrenia (pp. 179–200). Amsterdam: Excerpta Medica.

    Google Scholar 

  • Mednick, S. A. (1970). Breakdown in individuals at high risk for schizophrenia: Possible predispositional perinatal factors. Mental Hygiene, 54, 50–63.

    Google Scholar 

  • Mednick, S. A. (1974). Electrodermal recovery and psychopathology. In S. A. Mednick, F. Schulsinger, J. Higgins, & B. Bell (Eds.), Genetics, environment and psychopathology (pp. 135–146). Amsterdam: Elsevier.

    Google Scholar 

  • Mednick, S. A. (1978). Berkson’s fallacy and high-risk research. In L. C. Wynne, R. L. Cromwell, & S. Matthysse (Eds.), The nature of schizophrenia: New approaches to research and treatment (pp. 442–452). New York: Wiley.

    Google Scholar 

  • Mednick, S. A., & McNeil, T. F. (1968). Current methodology in research on the etiology of schizophrenia: Serious difficulties which suggest the use of the high-risk-group method. Psychological Bulletin, 70, 681–693.

    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 

  • Mednick, S. A., & Schulsinger, F. (1973). A learning theory of schizophrenia: Thirteen years later. In M. Hammer, K. Salzinger, & S. Sutton (Eds.), Psychopathology (pp. 343–360). New York: Wiley.

    Google Scholar 

  • Mednick, S. A., & Schulsinger, F. (1974). Studies of children at high risk for schizophrenia. In S. A. Mednick, F. Schulsinger, J. Higgins, & B. Bell (Eds.), Genetics, environment, and psychopathology (pp. 109–116). Amsterdam: Elsevier.

    Google Scholar 

  • Mednick, S. A., Schulsinger, F., Teasdale, T. W., Schulsinger, H., Venables, P. H., & Rock, D. R. (1978). Schizophrenia in high-risk children: Sex differences in predisposing factors. In G. Serban (Ed.), Cognitive defects in the developement of mental illness (pp. 169–197). New York: Brunner/Mazel.

    Google Scholar 

  • Melis, C., & van Boxtel, A. (2001). Differences in autonomic physiological responses between good and poor inductive reasoners. Biological Psychology, 58, 121–146.

    Article  PubMed  Google Scholar 

  • Meyers, M., & Smith, B. D. (1986). Hemispheric asymmetry and emotion: Effects of nonverbal affective stimuli. Biological Psychology, 22, 11–22.

    Article  PubMed  Google Scholar 

  • Meyers, M. B., & Smith, B. D. (1987). Cerebral processing of nonverbal affective stimuli: Differential effects of cognitive and affective sets on hemispheric asymmetry. Biological Psychology, 24, 67–84.

    Article  PubMed  Google Scholar 

  • Michaels, R. M. (1960). Tension responses of drivers generated on urban streets. Highway Research Board Bulletin, 271, 29–43.

    Google Scholar 

  • Michaels, R. M. (1962). Effect of expressway design on driver tension responses. Highway Research Board Bulletin, 330, 16–26.

    Google Scholar 

  • Miezejeski, C. M. (1978). Relationships between behavioral arousal and some components of autonomic arousal. Psychophysiology, 15, 417–421.

    Article  PubMed  Google Scholar 

  • Miller, N. E. (1969). Learning of visceral and glandular responses. Science, 163, 434–445.

    Article  PubMed  Google Scholar 

  • Miller, N. E. (1972). Learning of glandular and visceral responses: Postscript. In D. Singh & C. T. Morgan (Eds.), Current status of physiological psychology: Readings (pp. 228–250). Monterey: Brooks/Cole.

    Google Scholar 

  • Miller, R. M., & Coger, R. W. (1979). Skin conductance conditioning with dyshidrotic eczema patients. The British Journal of Dermatology, 101, 435–440.

    Article  PubMed  Google Scholar 

  • Miller, S., & Konorski, J. (1928). Sur une farme particulière des reflexes conditionels. Comptes Rendues Sociétée Biologique Paris, 99, 1155–1177.

    Google Scholar 

  • Miossec, Y., Catteau, M. C., Freixa i Baqué, E., & Roy, J.-C. (1985). Methodological problems in bilateral electrodermal research. International Journal of Psychophysiology, 2 , 247–256.

    Google Scholar 

  • Moan, E. R. (1979). GSR biofeedback assisted relaxation training and psychosomatic hives. Journal of Behavior Therapy and Experimental Psychiatry, 10, 157–158.

    Article  Google Scholar 

  • Mobascher, A., Brinkmeyer, J., Warbrick, T., Musso, F., Wittsack, H. J., Stoermer, R., et al. (2009). Fluctuations in electrodermal activity reveal variations in single trial brain responses to painful laser stimuli – A fMRI/EEG study. NeuroImage, 44, 1081–1092.

    Article  PubMed  Google Scholar 

  • Monat, A., Averill, J. R., & Lazarus, R. S. (1972). Anticipatory stress and coping reactions under various conditions of uncertainty. Journal of Personality and Social Psychology, 24, 237–253.

    Article  PubMed  Google Scholar 

  • Montagu, J. D. (1963). Habituation of the psycho-galvanic reflex during serial tests. Journal of Psychosomatic Research, 7, 199–214.

    Article  Google Scholar 

  • Morris, P. H., & Gale, A. (1993). Effects of situational demands on the direction of electrodermal activation during smoking. Addictive Behaviors, 18, 35–40.

    Article  PubMed  Google Scholar 

  • Morrow, L., Vrtunski, P. B., Kim, Y., & Boller, F. (1981). Arousal responses to emotional stimuli and laterality of lesion. Neuropsychologia, 19, 65–71.

    Article  PubMed  Google Scholar 

  • Mowrer, O. H. (1960). Learning theory and behavior. New York: Wiley.

    Book  Google Scholar 

  • Mundy-Castle, A. C., & McKiever, B. L. (1953). The psychophysiological significance of the galvanic skin response. Journal of Experimental Psychology, 46, 15–24.

    Article  PubMed  Google Scholar 

  • Munro, L. L., Dawson, M. E., Schell, A. M., & Sakai, L. M. (1987). Electrodermal lability and rapid vigilance decrement in a degraded stimulus continuous performance task. Journal of Psychophysiology, 1, 249–257.

    Google Scholar 

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

    Google Scholar 

  • Myrtek, M. (1984). Constitutional psychophysiology. Research in review. Orlando: Academic.

    Google Scholar 

  • Myslobodsky, M. S., & Rattok, J. (1975). Asymmetry of electrodermal activity in man. Bulletin of the Psychonomic Society, 6, 501–502.

    Google Scholar 

  • Myslobodsky, M. S., & Rattok, J. (1977). Bilateral electrodermal activity in waking man. Acta Psychologica, 41, 273–282.

    Article  PubMed  Google Scholar 

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

    Article  PubMed  Google Scholar 

  • National Research Council (2003). The polygraph and lie detection. Committee to Review the Scientific Evidence on the Polygraph. Division of Behavioral and Social Sciences and Education. Washington, DC: The National Academies Press.

    Google Scholar 

  • Naveteur, J., & Freixa i Baqué, E. (1987). Individual differences in electrodermal activity as a function of subjects’ anxiety. Personality and Individual Differences, 8, 615–626.

    Google Scholar 

  • Naveteur, J., Buisine, S., & Gruzelier, J. H. (2005). The influence of anxiety on electrodermal responses to distractors. International Journal of Psychophysiology, 56, 261–269.

    Article  PubMed  Google Scholar 

  • Naveteur, J., & Sequeira-Martinho, H. (1990). Reliability of bilateral differences in electrodermal activity. Biological Psychology, 31, 47–56.

    Article  Google Scholar 

  • Neary, R. S., & Zuckerman, M. (1976). Sensation seeking, trait and state anxiety, and the electrodermal orienting response. Psychophysiology, 13, 205–211.

    Article  PubMed  Google Scholar 

  • Nebylitsyn, V. D. (1972). Fundamental properties of the human nervous system. New York: Plenum.

    Google Scholar 

  • Nebylitsyn, V. D. (1973). Current problems in differential psychophysiology. Soviet Psychology, 11, 47–70.

    Google Scholar 

  • Neisser, U. (1967). Cognitive psychology. New York: Appleton.

    Google Scholar 

  • Netter, P. (1986). Einflußfaktoren auf die zentral-nervöse Wirkung von Beta- Rezeptorenblockern. In W. Janke & P. Netter (Eds.), Angst und Psychopharmaka (pp. 169–204). Stuttgart: Kohlhammer.

    Google Scholar 

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

    Article  PubMed  Google Scholar 

  • Nielsen, T. C., & Petersen, K. E. (1976). Electrodermal correlates of extraversion, trait anxiety and schizophrenism. Scandinavian Journal of Psychology, 17, 73–80.

    Article  PubMed  Google Scholar 

  • Niemelä, P. (1975). Effects of interrupting the process of preparation for film stress. Scandinavian Journal of Psychology, 16, 294–302.

    Article  Google Scholar 

  • Niepel, M. (2001). Independent manipulation of stimulus change and unexpectedness dissociates indices of the orienting response. Psychophysiology, 38, 84–91.

    Article  PubMed  Google Scholar 

  • Nikula, R. (1991). Psychological correlates of nonspecific skin conductance responses. Psychophysiology, 28, 86–90.

    Article  PubMed  Google Scholar 

  • Nomikos, M. S., Opton, E., Averill, J. R., & Lazarus, R. S. (1968). Surprise versus suspense in the production of stress reaction. Journal of Personality and Social Psychology, 8, 204–208.

    Article  PubMed  Google Scholar 

  • Norris, C. J., Larsen, J. T., & Cacioppo, J. T. (2007). Neuroticism is associated with larger and more prolonged electrodermal responses to emotionally evocative pictures. Psychophysiology, 44, 823–826.

    Article  PubMed  Google Scholar 

  • Nuechterlein, K. H. (1987). Vulnerability models for schizophrenia: State of the art. In H. Häfner, W. F. Gattaz, & W. Janzarik (Eds.), Search for the causes of schizophrenia (pp. 297–316). Heidelberg: Springer.

    Chapter  Google Scholar 

  • Nuechterlein, K. H., & Dawson, M. E. (1984). A heuristic vulnerability/stress model of schizophrenic episodes. Schizophrenia Bulletin, 10, 300–312.

    PubMed  Google Scholar 

  • Nuechterlein, K. H., Edell, W. S., Norris, M., & Dawson, M. E. (1986). Attentional vulnerability indicators, thought disorders, and negative symptoms in schizophrenia. Schizophrenia Bulletin, 12, 408–426.

    PubMed  Google Scholar 

  • O’Gorman, J. G. (1974). A comment on Koriat, Averill, and Malmstrom’s “Individual differences in habituation”. Journal of Research in Personality, 8, 198–202.

    Article  Google Scholar 

  • O’Gorman, J. G. (1977). Individual differences in habituation of human physiological responses: A review of theory, method, and findings in the study of personality correlates in non-clinical populations. Biological Psychology, 5, 257–319.

    Article  PubMed  Google Scholar 

  • O’Gorman, J. G. (1978). Method of recording: A neglected factor in the controversy over the bimodality of electrodermal responsiveness in schizophrenic samples. Schizophrenia Bulletin, 4, 150–152.

    PubMed  Google Scholar 

  • O’Gorman, J. G. (1979). The orienting reflex: Novelty or significance detector? Psychophysiology, 16, 253–262.

    Article  PubMed  Google Scholar 

  • O’Gorman, J. G., & Lloyd, J. E. M. (1988). Electrodermal lability and dichotic listening. Psychophysiology, 25, 538–546.

    Article  PubMed  Google Scholar 

  • O’Toole, D., Yuille, J. C., Patrick, C. J., & Iacono, W. G. (1994). Alcohol and the physiological detection of deception: Arousal and memory influences. Psychophysiology, 31, 253–263.

    Article  PubMed  Google Scholar 

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

    Article  PubMed  Google Scholar 

  • Obrist, P. A. (1976). The cardiovascular-behavioral interaction – As it appears today. Psychophysiology, 13, 95–107.

    Article  PubMed  Google Scholar 

  • Obrist, P. A., Black, F. W., Brener, J., & DiCara, L. W. (Eds.). (1974). Cardiovascular psychophysiology. Chicago: Aldine.

    Google Scholar 

  • Offe, H., & Offe, S. (2007). The comparison question test: Does it work and if so how? Law and Human Behavior, 31, 291–303.

    Article  PubMed  Google Scholar 

  • Ogawa, T. (1984). Regional differences in sweating activity. In J. R. S. Hales (Ed.), Thermal physiology (pp. 229–234). New York: Raven.

    Google Scholar 

  • Öhman, A. (1971). Differentiation of conditioned and orienting response components in electrodermal conditioning. Psychophysiology, 8, 7–22.

    Article  PubMed  Google Scholar 

  • Öhman, A. (1979). The orienting response, attention, and learning: An information-processing perspective. In H. D. Kimmel, E. H. van Olst, & J. F. Orlebeke (Eds.), The orienting reflex in humans (pp. 443–471). Hillsdale, NJ: Erlbaum.

    Google Scholar 

  • Öhman, A. (1981). Electrodermal activity and vulnerability to schizophrenia: A review. Biological Psychology, 12, 87–145.

    Article  PubMed  Google Scholar 

  • Öhman, A. (1983). The orienting response during Pavlovian conditioning. In D. Siddle (Ed.), Orienting and habituation: Perspectives in human research (pp. 315–369). Chichester: Wiley.

    Google Scholar 

  • Öhman, A. (1993). Stimulus prepotency and fear learning: Data and theory. In N. Birbaumer & A. Öhman (Eds.), The structur of emotion (pp. 218–239). Seattle: Hogrefe & Huber.

    Google Scholar 

  • Öhman, A., & Bohlin, G. (1973). The relationship between spontaneous and stimulus-correlated electrodermal responses in simple and discriminative conditioning paradigms. Psychophysiology, 10, 589–600.

    Article  PubMed  Google Scholar 

  • Öhman, A., Dimberg, U., & Esteves, F. (1989). Preattentive activation of aversive emotions. In T. Archer & L.-G. Nilsson (Eds.), Aversion, avoidance, and anxiety (pp. 169–193). Hillsdale, NJ: Erlbaum.

    Google Scholar 

  • Öhman, A., Erikson, A., & Olofsson, C. (1975). One-trial learning and superior resistance to extinction of autonomic responses conditioned to potentially phobic stimuli. Journal of Comparative and Physiological Psychology, 88, 619–627.

    Article  PubMed  Google Scholar 

  • Öhman, A., Eriksson, A., Fredrikson, M., Hugdahl, K., & Olofsson, C. (1974). Habituation of the electrodermal orienting reaction to potentially phobic and supposedly neutral stimuli in normal human subjects. Biological Psychology, 2, 85–93.

    Article  PubMed  Google Scholar 

  • Öhman, A., Erixon, G., & Löfberg, G. (1975). Phobias and preparedness: Phobic versus neutral pictures as conditioned stimuli for human autonomic responses. Journal of Abnormal Psychology, 84, 41–45.

    Article  PubMed  Google Scholar 

  • Öhman, A., & Hultman, C. M. (1998). Electrodermal activity and obstetric complications in schizophrenia. Journal of Abnormal Psychology, 107, 228–237.

    Article  PubMed  Google Scholar 

  • Öhman, A., & Soares, J. J. (1994). “Unconscious anxiety”: Phobic responses to masked stimuli. Journal of Abnormal Psychology, 103, 231–240.

    Article  PubMed  Google Scholar 

  • Olbrich, R. (1989). Electrodermal activity and its relevance to vulnerability research in schizophrenics. The British Journal of Psychiatry, 155, 40–45.

    Google Scholar 

  • Olbrich, R. (1990). The contributions of psychophysiology to vulnerability models. In H. Häfner & W. F. Gattaz (Eds.), Search for the causes of schizophrenia (Vol. 1, pp. 192–204). Berlin: Springer.

    Chapter  Google Scholar 

  • Olbrich, R., & Mussgay, L. (1987). Spontaneous fluctuations of electrical skin conductance and the actual clinical state in schizophrenics. Psychopathology, 20, 18–22.

    Article  PubMed  Google Scholar 

  • Olds, J., & Olds, M. E. (1965). Drives, rewards, and the brain. In T. M. Newcomb (Ed.), New directions in psychology II (pp. 327–410). New York: Holt, Rinehart, & Winston.

    Google Scholar 

  • Orlebeke, J. F., & van Olst, E. H. (1968). Learning and performance as a function of CS-intensity in a delayed GSR conditioning situation. Journal of Experimental Psychology, 77, 483–487.

    Article  PubMed  Google Scholar 

  • Orr, S. P., & Lanzetta, J. T. (1980). Facial expressions of emotion as conditioned stimuli for human autonomic responses. Journal of Personality and Social Psychology, 38, 278–282.

    Article  PubMed  Google Scholar 

  • Orr, S. P., Milad, M. R., Metzger, L. J., Lasko, N. B., Gilbertson, M. W., & Pitman, R. K. (2006). Effects of beta blockade, PTSD diagnosis, and explicit threat on the extinction and retention of an aversively conditioned response. Biological Psychology, 73, 262–271.

    Article  PubMed  Google Scholar 

  • Oscar-Berman, M., & Gade, A. (1979). Electrodermal measures of arousal in humans with cortical or subcortical brain damage. In H. D. Kimmel, E. H. van Olst, & J. F. Orlebeke (Eds.), The orienting reflex in humans (pp. 665–676). Hillsdale, NJ: Erlbaum.

    Google Scholar 

  • Overall, J. E., & Woodward, J. A. (1977). Nonrandom assignment and the analysis of covariance. Psychological Bulletin, 84, 588–594.

    Article  Google Scholar 

  • Overmier, J. B. (1985). Toward a reanalysis of the causal structure of the learned helplessness syndrome. In F. R. Brush & J. B. Overmier (Eds.), Affect, conditioning, and cognition: Essays on the determinants of behavior (pp. 211–227). Hillsdale, NJ: Erlbaum.

    Google Scholar 

  • Packer, J. S., & Siddle, D. A. T. (1989). Stimulus miscuing, electrodermal activity, and the allocation of processing resources. Psychophysiology, 26, 192–200.

    Article  PubMed  Google Scholar 

  • Page, M., & Robson, A. (2007). Galvanic skin responses from asking stressful questions. The British Journal of Nursing, 16, 622–627.

    Google Scholar 

  • Panksepp, J. (1982). Toward a general psychobiological theory of emotions. The Behavioral and Brain Sciences, 5, 407–467.

    Article  Google Scholar 

  • Panksepp, J. (1986). The anatomy of emotions. In R. Plutchik & H. Kellermann (Eds.), Emotion: Theory research and experience (pp. 91–124). New York: Academic.

    Google Scholar 

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

    Google Scholar 

  • Patterson, T. (1976). Skin conductance recovery and pupillometrics in chronic schizophrenia. Psychophysiology, 13, 189–195.

    Article  PubMed  Google Scholar 

  • Patterson, T., & Venables, P. H. (1981). Bilateral skin conductance and the pupillary light-dark reflex: Manipulation by chlorpromazine, haloperidol, scopolamine, and placebo. Psychopharmacology, 73, 63–69.

    Article  PubMed  Google Scholar 

  • Pavlov, I. P. (1927). Conditioned reflexes. An investigation of the physiological activity of the cerebral cortex. New York: Oxford University Press.

    Google Scholar 

  • Peeke, S. C., & Grings, W. W. (1968). Magnitude of UCR as a function of variability in the CS-UCS relationship. Journal of Experimental Psychology, 77, 64–69.

    Article  PubMed  Google Scholar 

  • Pelachaud, C., & Bilvi, M. (2003). Computational model of believable conversational agents. In M.-P. Huget (Ed.), Communications in multiagent systems (pp. 300–317). Berlin: Springer.

    Chapter  Google Scholar 

  • Pennebaker, J. W., & Chew, C. H. (1985). Behavioral inhibition and electrodermal activity during deception. Journal of Personality and Social Psychology, 49, 1427–1433.

    Article  PubMed  Google Scholar 

  • Peper, M., Karcher, S., Wohlfart, R., Reinshagen, G., & LeDoux, J. E. (2001). Aversive learning in patients with unilateral lesions of the amygdala and hippocampus. Biological Psychology, 58, 1–23.

    Article  PubMed  Google Scholar 

  • Perry, W., Felger, T., & Braff, D. (1998). The relationship between skin conductance hyporesponsivity and perseverations in schizophrenia patients. Biological Psychiatry, 44, 459–465.

    Article  PubMed  Google Scholar 

  • Peters, T. (1974). Mentale Beanspruchung von Büroangestellten im Schreibdienst und bei Vorzimmertätigkeit. Zentralblatt für Arbeitsmedizin und Arbeitsschutz, 24, 197–207.

    PubMed  Google Scholar 

  • Petrinovich, L. (1973). A species-meaningful analysis of habituation. In H. V. S. Peeke & M. J. Herz (Eds.), habituation (Behavioral studies, Vol. 1, pp. 141–162). New York: Academic.

    Google Scholar 

  • Phillips, K. C., Evans, P. D., & Fearn, J. M. (1986). Heart rate and skin conductance correlates of monitoring or distraction as strategies for “coping”. In D. Papakostopoulos, S. Butler, & I. Martin (Eds.), Clinical and experimental neuropsychophysiology (pp. 486–499). Dover: Croom Helm.

    Google Scholar 

  • Picard, R. W., & Healy, J. (1997). Affective wearables. Personal and Ubiquitous Computing, 1, 231–240.

    Article  Google Scholar 

  • Pineles, S. L., Orr, M. R., & Orr, S. P. (2009). An alternative scoring method for skin conductance responding in a differential fear conditioning paradigm with a long-duration conditioned stimulus. Psychophysiology, 46, 984–995.

    Article  PubMed  Google Scholar 

  • Pinkus, H. (1952). Examination of the epidermis by the strip method: II. Biometric data on regeneration of the human epidermis. Journal of Investigative Dermatology, 19, 431–447.

    Article  PubMed  Google Scholar 

  • Pivik, R. T. (1978). Tonic states and phasic events in relation to sleep mentation. In M. A. Arkin, J. S. Antrobus, & S. Ellman (Eds.), The mind in sleep: Physiology and psychophysiology (pp. 245–271). New York: Wiley.

    Google Scholar 

  • Pivik, R. T. (2007). Sleep and dreaming. In J. T. Cacioppo, L. G. Tassinary, & G. Berntson (Eds.), Handbook of psychophysiology (pp. 633–662). New York: Cambridge University Press.

    Google Scholar 

  • Plouffe, L., & Stelmack, R. M. (1986). Sensation-seeking and the electrodermal orienting response in young and elderly females. Personality and Individual Differences, 7, 119–120.

    Article  Google Scholar 

  • Plutchik, R. (1980). Emotion – A psychoevolutionary synthesis. New York: Harper & Row.

    Google Scholar 

  • Pollatos, O., Schubö, A., Herbert, B. M., Matthias, E., & Schandry, R. (2008). Deficits in early emotional reactivity in alexithymia. Psychophysiology, 45, 839–846.

    PubMed  Google Scholar 

  • Porges, S. W., & Fox, N. A. (1986). Developmental psychophysiology. In M. G. H. Coles, E. Donchin, & S. W. Porges (Eds.), Psychophysiology: Systems, processes, and applications (pp. 611–625). Amsterdam: Elsevier.

    Google Scholar 

  • Posner, M. I. (1975). Psychobiology of attention. In M. S. Gazzaniga & C. Blakemore (Eds.), Handbook of psychobiology (pp. 441–480). New York: Academic.

    Google Scholar 

  • Prentky, A., Salzman, L. F., & Klein, R. H. (1981). Habituation and conditioning of skin conductance responses in children at risk. Schizophrenia Bulletin, 7, 281–291.

    PubMed  Google Scholar 

  • Preston, B. (1969). Insurance classifications and drivers’ galvanic skin response. Ergonomics, 12, 437–446.

    Article  PubMed  Google Scholar 

  • Pribram, K. H. (1980). The biology of emotions and other feelings. In R. Plutchik & H. Kellerman (Eds.), Emotion: Theory, research, and experience (Theories of emotion, Vol. 1, pp. 245–269). New York: Academic.

    Google Scholar 

  • Pribram, K. H., & McGuinness, D. (1975). Arousal, activation, and effort in the control of attention. Psychological Review, 82, 116–149.

    Article  PubMed  Google Scholar 

  • Pribram, K. H., & McGuinness, D. (1976). Arousal, Aktivierung und Anstrengung: Gesonderte neurale Systeme. Zeitschrift für Psychologie, 184, 382–403.

    PubMed  Google Scholar 

  • Prior, M. G., Cumming, G., & Hendy, J. (1984). Recognition of abstract and concrete words in a dichotic listening paradigm. Cortex, 20, 149–157.

    PubMed  Google Scholar 

  • Prokasy, W. F., & Ebel, H. C. (1967). Three components of the classically conditioned GSR in human subjects. Journal of Experimental Psychology, 73, 247–256.

    Article  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 

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

    Google Scholar 

  • Rabavilas, A. D. (1987). Electrodermal activity in low and high alexithymia neurotic patients. Psychotherapy and Psychosomatics, 47, 101–104.

    Article  PubMed  Google Scholar 

  • Raine, A. (2002). Annotation: The role of prefrontal deficits, low autonomic arousal, and early health factors in the development of antisocial and aggressive behavior in children. Journal of Child Psychology and Psychiatry, 43, 417–434.

    Article  PubMed  Google Scholar 

  • Raine, A., Bihrle, S., Venables, P. H., Mednick, S. A., & Pollock, V. (1999). Skin-conductance orienting deficits and increased alcoholism in schizotypal criminals. Journal of Abnormal Psychology, 108, 299–306.

    Article  PubMed  Google Scholar 

  • Raine, A., Lencz, T., Bihrle, S., LaCasse, L., & Colletti, P. (2000). Reduced prefrontal gray matter volume and reduced autonomic activity in antisocial personality disorder. Archives of General Psychiatry, 57, 119–127.

    Article  PubMed  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. (1981). Classical conditioning and socialization – A biosocial interaction. Personality and Individual Differences, 2, 273–283.

    Article  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 

  • Raine, A., Venables, P. H., & Williams, M. (1990a). Relationship between central and autonomic measures of arousal at age 15 years and criminality at age 24 years. Archives of General Psychiatry, 47, 1003–1007.

    PubMed  Google Scholar 

  • Raine, A., Venables, P. H., & Williams, M. (1990b). Autonomic orienting responses in 15-year-old male subjects and criminal behavior at age 24. The American Journal of Psychiatry, 147, 933–937.

    PubMed  Google Scholar 

  • Raine, A., Venables, P. H., & Williams, M. (1995). High autonomic arousal and electrodermal orienting at age 15 years as protective factors against criminal behavior at age 29 years. The American Journal of Psychiatry, 152, 1595–1600.

    PubMed  Google Scholar 

  • Rajamanickam, M., & Gnanaguru, K. (1981). Physiological correlates of personality. Psychological Studies, 26, 41–43.

    Google Scholar 

  • Rakov, G. V., & Fadeev, Y. A. (1986). Assessment of emotional stress during work activity by system analysis of the galvanic skin reflex. Human Physiology, 11, 215–220.

    Google Scholar 

  • Rappaport, H., & Katkin, E. S. (1972). Relationships among manifest anxiety, response to stress, and the perception of autonomic activity. Journal of Consulting and Clinical Psychology, 38, 219–224.

    Article  PubMed  Google Scholar 

  • Raskin, D. C. (1973). Attention and arousal. In W. F. Prokasy & D. C. Raskin (Eds.), Electrodermal activity in psychological research (pp. 125–155). New York: Academic.

    Google Scholar 

  • Raskin, M. (1975). Decreased skin conductance response habituation in chronically anxious patients. Biological Psychology, 2, 309–319.

    Article  PubMed  Google Scholar 

  • Raskin, D. C. (1979). Orienting and defensive reflexes in the detection of deception. In H. D. Kimmel, E. H. van Olst, & J. F. Orlebeke (Eds.), The orienting reflex in humans (pp. 587–605). Hillsdale, NJ: Erlbaum.

    Google Scholar 

  • Raskin, D. C., & Honts, C. R. (2002). The Comparison Question Test. In M. Kleiner (Ed.), Handbook of polygraph testing (pp. 1–48). San Diego, CA: Academic.

    Google Scholar 

  • Raskin, D. C., Kotses, H., & Bever, J. (1969). Autonomic indicators of orienting and defensive reflexes. Journal of Experimental Psychology, 3, 423–433.

    Article  Google Scholar 

  • Raskin, D. C., & Podlesny, J. A. (1979). Truth and deception: A reply to Lykken. Psychological Bulletin, 86, 54–59.

    Article  PubMed  Google Scholar 

  • Reeves, B., Lang, A., Kim, E. Y., & Tatar, D. (1999). The effects of screen size and message content on attention and arousal. Media Psychology, 1, 49–67.

    Article  Google Scholar 

  • Rescorla, R. A. (1967). Pavlovian conditioning and its proper control procedures. Psychological Review, 74, 71–80.

    Article  PubMed  Google Scholar 

  • Rescorla, R. A. (1969). Pavlovian conditioned inhibition. Psychological Bulletin, 72, 77–94.

    Article  Google Scholar 

  • Rescorla, R. A. (1988). Pavlovian conditioning: It’s not what you think it is. American Psychologist, 43, 151–160.

    Article  PubMed  Google Scholar 

  • Rice, D. G. (1966). Operant conditioning and associated electromyogram responses. Journal of Experimental Psychology, 71, 908–912.

    Article  PubMed  Google Scholar 

  • Richter, P., Wagner, T., Heger, R., & Weise, G. (1998). Psychophysiological analysis of mental load during driving on rural roads – A quasi-experimental field study. Ergonomics, 41, 593–609.

    Article  PubMed  Google Scholar 

  • Rickels, K. (1978). Use of antianxiety agents in anxious outpatients. Psychopharmacology, 58, 1–17.

    Article  PubMed  Google Scholar 

  • Ridgeway, D., & Hare, R. D. (1981). Sensation seeking and psychophysiological responses to auditory stimulation. Psychophysiology, 18, 613–618.

    Article  PubMed  Google Scholar 

  • Riley, L. H., & Richter, C. P. (1975). Uses of the electrical skin resistance method in the study of patients with neck and upper extremity pain. The Johns Hopkins Medical Journal, 137, 69–74.

    PubMed  Google Scholar 

  • Rippon, G. (1990). Individual differences in electrodermal and electroencephalographic asymmetries. International Journal of Psychophysiology, 8, 309–320.

    Article  PubMed  Google Scholar 

  • Ritchie, B. H. (1953). The circumnavigation of cognition. Psychological Review, 60, 216–221.

    Article  PubMed  Google Scholar 

  • Rizzolatti, G., & Buchtel, H. A. (1977). Hemispheric superiority in reaction time to faces: A sex difference. Cortex, 13, 300–305.

    PubMed  Google Scholar 

  • Roberts, L. E. (1974). Comparative psychophysiology of the electrodermal and cardiac control systems. In P. A. Obrist, A. H. Black, J. Brener, & L. V. DiCara (Eds.), Cardiovascular psychophysiology (pp. 163–189). Chicago: Aldine.

    Google Scholar 

  • Roberts, L. E. (1977). The role of exteroceptive feedback in learned electrodermal and cardiac control: Some attractions of and problems with discrimination theory. In J. Beatty & H. Legewie (Eds.), Biofeedback and behavior (pp. 261–280). New York: Plenum.

    Google Scholar 

  • Roberts, L. E., Lacroix, J. M., & Wright, M. (1974). Comparative studies of operant electrodermal and heart rate conditioning in curarized rats. In P. A. Obrist, A. H. Black, J. Brener, & L. V. DiCara (Eds.), Cardiovascular psychophysiology (pp. 332–352). Chicago: Aldine.

    Google Scholar 

  • Rodriguez, S., Fernandez, M. C., Cependa-Benito, A., & Vila, J. (2005). Subjective and physiological reactivity to chocolate images in high and low chocolate cravers. Biological Psychology, 70, 9–18.

    Article  PubMed  Google Scholar 

  • Roedema, T. M., & Simons, R. F. (1999). Emotion-processing deficit in alexithymia. Psychophysiology, 36, 279–387.

    Article  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 

  • Rommelspacher, H. (1981). The beta-carbolines (harmanes): A new class of endogenous compounds: Their relevance for the pathogenesis and treatment of psychiatric and neurological diseases. Pharmacopsychiatry, 14, 117–125.

    Article  Google Scholar 

  • Rosenman, R. H., Friedman, M., Straus, R., Wurm, M., Jenkins, C. D., & Messinger, H. B. (1966). Coronary heart disease in the Western Collaborative Group Study. A follow-up experience of two years. Journal of the American Medical Association, 195, 130–136.

    Article  Google Scholar 

  • Rotenberg, V. S., & Vedenyapin, A. B. (1985). GSR as reflection of decision-making under conditions of delay. The Pavlovian Journal of Biological Science, 20, 11–14.

    PubMed  Google Scholar 

  • Roth, W. T., Goodale, J., & Pfefferbaum, A. (1991). Auditory event-related potentials and electrodermal activity in medicated and unmedicated schizophrenics. Biological Psychiatry, 29, 585–599.

    Article  PubMed  Google Scholar 

  • Roth, W. T., Wilhelm, F. H., & Trabert, W. (1998). Autonomic instability during relaxation in panic disorder. Psychiatry Research, 80, 155–164.

    Article  PubMed  Google Scholar 

  • Routtenberg, A. (1968). The two-arousal hypothesis: Reticular formation and limbic system. Psychological Review, 75, 51–80.

    Article  PubMed  Google Scholar 

  • Routtenberg, A. (1971). Stimulus processing and response execution: A neurobehavioral theory. Physiology & Behavior, 6, 589–596.

    Article  Google Scholar 

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

    Google Scholar 

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

    Chapter  Google Scholar 

  • Rubens, R., & Lapidus, L. B. (1978). Schizophrenic patterns of arousal and stimulus barrier functioning. Journal of Abnormal Psychology, 87, 199–211.

    Article  PubMed  Google Scholar 

  • Rushby, J. A., & Barry, R. J. (2007). Event-related potential correlates of phasic and tonic measures of the orienting reflex. Biological Psychology, 75, 248–259.

    Article  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 

  • Saari, A., Tolonen, U., Pääkkö, E., Suominen, K., Pyhtinen, J., Sotaniemi, K. A., et al. (2008). Sympathetic skin responses in multiple sclerosis. Acta Neurologica Scandinavica, 118, 226–231.

    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 

  • Salzman, L. F., & Klein, R. H. (1978). Habituation and conditioning of electrodermal responses in high-risk children. Schizophrenia Bulletin, 4, 210–222.

    PubMed  Google Scholar 

  • Sartory, G. (1983). The orienting response and psychopathology: Anxiety and phobias. In D. Siddle (Ed.), Orienting and habituation: Perspectives in human research (pp. 449–474). Chichester: Wiley.

    Google Scholar 

  • Satterfield, J. H., & Dawson, M. E. (1971). Electrodermal correlates of hyperactivity in children. Psychophysiology, 8, 191–197.

    Article  PubMed  Google Scholar 

  • Scarpa, A., & Raine, A. (1997). Psychophysiology of anger and violent behavior. The Psychiatric Clinics of North America, 20, 375–394.

    Article  PubMed  Google Scholar 

  • Scerbo, M. W., Freeman, F. G., & Mikulka, P. J. (2000). A biocybernetic system for adaptive automation. In R. W. Backs & W. Boucsein (Eds.), Engineering psychophysiology. Issues and applications (pp. 241–253). Mahwah, NY: Lawrence Erlbaum.

    Google Scholar 

  • Schachter, S., & Singer, J. E. (1962). Cognitive, social and physiological determinants of emotional state. Psychological Review, 69, 379–399.

    Article  PubMed  Google Scholar 

  • Schaefer, F., Kuhmann, W., Boucsein, W., & Alexander, J. (1986). Beanspruchung durch Bildschirmtätigkeit bei experimentell variierten Systemresponsezeiten. Zeitschrift für Arbeitswissenschaft, 40(12 NF), 31–38.

    Google Scholar 

  • Schaefer, F., Schäfer, R., & Boucsein, W. (2000). Auswirkungen von Prozesslaufzeit und Prozessindikatoren beim Multi-Tasking auf Arbeitsstrategie und Beanspruchung des Benutzers. Zeitschrift für Arbeitswissenschaft, 54, 267–275.

    Google Scholar 

  • Schandry, R. (1978). Habituation psychophysiologischer Größen in Abhängigkeit von der Reizintensität. München: Minerva.

    Google Scholar 

  • Schell, A. M., Dawson, M. E., & Filion, D. I. (1988). Psychophysiological correlates of electrodermal lability. Psychophysiology, 25, 619–632.

    Article  PubMed  Google Scholar 

  • Schell, A. M., Dawson, M. E., Nuechterlein, K. H., Subotnik, K. L., & Ventura, J. (2002). The temporal stability of electrodermal variables over a one-year period in patients with recent-onset schizophrenia and in normal subjects. Psychophysiology, 39, 124–132.

    Article  PubMed  Google Scholar 

  • Schell, A. M., Dawson, M. E., Rissling, A., Ventura, J., Subotnik, K. L., Gitlin, M. J., et al. (2005). Electrodermal predictors of functional outcome and negative symptoms in schizophrenia. Psychophysiology, 42, 483–492.

    Article  PubMed  Google Scholar 

  • Schiffmann, K., & Furedy, J. J. (1972). Failures of contingency and cognitive factors to affect long-interval differential Pavlovian autonomic conditioning. Journal of Experimental Psychology, 96, 215–218.

    Article  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 

  • Schmidt, S., Schneider, R., Binder, M., Bürkle, D., & Walach, H. (2001). Investigating methodological issues in EDA-DMILS: Results from a pilot study. The Journal of Parapsychology, 65, 59–82.

    Google Scholar 

  • Schmidt, S., & Walach, H. (2000). Electrodermal activity (EDA) – State-of-the-art measurement and techniques for parapsychological purposes. The Journal of Parapsychology, 64, 139–164.

    Google Scholar 

  • Schnur, D. B., Bernstein, A. S., Mukherjee, S., Loh, J., Degreef, G., & Reidel, J. (1989). The autonomic orienting response and CT scan findings in schizophrenia. Schizophrenia Research, 2, 449–455.

    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 

  • Schulter, G., & Papousek, I. (1998). Bilateral electrodermal activity: Relationsships to state and trait characteristics of hemisphere asymmetry. International Journal of Psychophysiology, 31, 1–12.

    Article  PubMed  Google Scholar 

  • Schuri, U., & von Cramon, D. (1979). Autonomic responses to meaningful and non-meaningful auditory stimuli in coma. Archiv für Psychiatrie und Nervenkrankheiten, 227, 143–149.

    PubMed  Google Scholar 

  • Schuri, U., & von Cramon, D. (1980). Autonomic and behavioral responses in coma due to drug overdose. Psychophysiology, 17, 253–258.

    Article  PubMed  Google Scholar 

  • Schuri, U., & von Cramon, D. (1981). Electrodermal responses to auditory stimuli with different significance in neurological patients. Psychophysiology, 18, 248–251.

    Article  PubMed  Google Scholar 

  • Schuri, U., & von Cramon, D. (1982). Electrodermal response patterns in neurological patients with disturbed vigilance. Behavioural Brain Research, 4, 95–102.

    Article  PubMed  Google Scholar 

  • Schwalen, S., Altermann, A., Jörg, J., Berg, K., & Cramer, B. M. (1996). Bilateral suppression of the sympathetic nervous system in hemispheric brain infarction. Journal of Neurology, 243, 157–160.

    Article  PubMed  Google Scholar 

  • Schwartz, G. E. (1986). Emotion and psychophysiological organization: A systems approach. In M. G. H. Coles, E. Donchin, & S. W. Porges (Eds.), Psychophysiology: Systems, processes, and applications (pp. 354–377). Amsterdam: Elsevier.

    Google Scholar 

  • Schwartz, G. E., & Shapiro, D. (1973). Social Psychophysiology. In W. F. Prokasy & D. C. Raskin (Eds.), Electrodermal activity in psychological research (pp. 377–416). New York: Academic.

    Google Scholar 

  • Segal, E. M., & Lachman, R. (1972). Complex behavior or higher mental process: Is there a paradigm shift? American Psychologist, 27, 45–55.

    Article  Google Scholar 

  • Seligman, M. E. P. (1969). Control group and conditioning: A comment on operationism. Psychological Review, 76, 484–491.

    Article  Google Scholar 

  • Seligman, M. E. P. (1971). Phobias and preparedness. Behavior Therapy, 2, 307–320.

    Article  Google Scholar 

  • Seligman, L. (1975). Skin potential as an indicator of emotion. Journal of Counseling Psychology, 22, 489–493.

    Article  Google Scholar 

  • Selye, H. (1976). The stress of life. New York: McGraw-Hill.

    Google Scholar 

  • Sequeira, H., Hot, P., Silvert, L., & Delplanque, S. (2009). Electrical autonomic correlates of emotion. International Journal of Psychophysiology, 71, 50–56.

    Article  PubMed  Google Scholar 

  • Shahani, B., Halperin, J. J., Boulu, P., & Cohen, J. (1984). Sympathetic skin response – a method of assessing unmyelinated axon dysfunction in peripheral neuropathies. Journal of Neurology, 47, 536–542.

    Google Scholar 

  • Shapiro, D. (1977). A monologue on biofeedback and psychophysiology. Psychophysiology, 14, 213–227.

    Article  PubMed  Google Scholar 

  • Sharpless, D., & Jasper, H. (1956). Habituation of the arousal reaction. Brain, 79, 655–681.

    Article  PubMed  Google Scholar 

  • Shneiderman, B. (1992). Response time and display rate. In B. Shneiderman (Ed.), Designing the user interface (pp. 277–301). Reading, MA: Addison-Wesley.

    Google Scholar 

  • Siddle, D. A. T. (1972). Vigilance decrement and speed of habituation of the GSR component of the orienting response. British Journal of Psychology, 63, 191–194.

    Article  PubMed  Google Scholar 

  • Siddle, D. A. T. (1977). Electrodermal activity and psychopathy. In S. A. Mednick & K. O. Christiansen (Eds.), Biosocial bases of criminal behavior (pp. 199–211). New York: Gardener.

    Google Scholar 

  • Siddle, D. (Ed.). (1983). Orienting and habituation: Perspectives in human research. Chichester: Wiley.

    Google Scholar 

  • Siddle, D. A. T. (1985). Effects of stimulus omission and stimulus change on dishabituation of the skin conductance response. Journal of Experimental Psychology. Learning, Memory, and Cognition, 11, 206–216.

    Article  PubMed  Google Scholar 

  • Siddle, D. A. T. (1991). Orienting, habituation, and resource allocation: An associative analysis. Psychophysiology, 28, 245–259.

    Article  PubMed  Google Scholar 

  • Siddle, D. A. T., & Heron, P. A. (1976). Reliability of electrodermal habituation measures under two conditions of stimulus intensity. Journal of Research in Personality, 10, 195–200.

    Article  Google Scholar 

  • Siddle, D. A. T., Lipp, O. V., & Dall, P. J. (1996). Effects of intermodality change and number of training trials on electodermal orienting and on the allocation of processing resources. Biological Psychology, 43, 57–67.

    Article  PubMed  Google Scholar 

  • Siddle, D. A. T., O’Gorman, J. G., & Wood, L. (1979). Effects of electrodermal lability and stimulus significance on electrodermal response amplitude to stimulus change. Psychophysiology, 16, 520–527.

    Article  PubMed  Google Scholar 

  • Siddle, D. A. T., & Packer, J. (1987). Stimulus omission and dishabituation of the electrodermal orienting response: The allocation of processing resources. Psychophysiology, 24, 181–190.

    Article  PubMed  Google Scholar 

  • Siddle, D. A. T., & Remington, B. (1987). Latent inhibition and human Pavlovian conditioning: Research and relevance. In G. Davey (Ed.), Conditioning in humans (pp. 115–146). Chichester: Wiley.

    Google Scholar 

  • Siddle, D. A. T., Remington, B., Kuiack, M., & Haines, E. (1983). Stimulus omission and dishabituation of the skin conductance response. Psychophysiology, 20, 136–145.

    Article  PubMed  Google Scholar 

  • Siddle, D., Stephenson, D., & Spinks, J. A. (1983). Elicitation and habituation of the orienting response. In D. Siddle (Ed.), Orienting and habituation: Perspectives in human research (pp. 109–182). Chichester: Wiley.

    Google Scholar 

  • Siddle, D. A. T., Turpin, G., Spinks, J. A., & Stephenson, D. (1980). Peripheral measures. In H. M. van Praag, M. H. Lader, O. J. Rafaelsen, & E. J. Sachar (Eds.), Handbook of biological psychiatry: Part 2. Brain mechanisms and abnormal behavior – Psychophysiology (pp. 45–78). New York: Dekker.

    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 

  • Silvert, L., Delplanque, S., Bouwalerh, H., Verpoort, C., & Sequeira, H. (2004). Autonomic responding to aversive words without conscious valence discrimination. International Journal of Psychophysiology, 53, 135–145.

    Article  PubMed  Google Scholar 

  • Silvestrini, N., & Gendolla, G. H. E. (2007). Mood effects on autonomic activity in mood regulation. Psychophysiology, 44, 650–659.

    Article  PubMed  Google Scholar 

  • Simons, R. F., Detenber, B. H., Roedema, T. M., & Reiss, J. E. (1999). Emotion processing in three systems: The medium and the message. Psychophysiology, 36, 619–627.

    Article  PubMed  Google Scholar 

  • Siniatchkin, M., Gerber, W.-D., Kropp, P., Voznesenskaya, T., & Vein, A. M. (2000). Are the periodic changes of neurophysiological parameters during the pain-free interval in migraine related to abnormal orienting activity? Cephalalgia, 20, 20–29.

    Article  PubMed  Google Scholar 

  • Smith, B. D., Gatchel, R. J., Korman, M., & Satter, S. (1979). EEG and autonomic responding to verbal, spatial and emotionally arousing tasks: Differences among adults, adolescents and inhalant abusers. Biological Psychology, 9, 189–200.

    Article  PubMed  Google Scholar 

  • Smith, B. D., Ketterer, M. W., & Concannon, M. (1981). Bilateral electrodermal activity as a function of hemispheric stimulation, hand preference, sex, and familial handedness. Biological Psychology, 12, 1–11.

    Article  PubMed  Google Scholar 

  • Smith, B. D., Kline, R., & Meyers, M. (1990). The differential hemisphere processing of emotion: A comparative analysis in strongly-lateralized sinistrals and dextrals. International Journal of Neuroscience, 50, 59–71.

    Article  PubMed  Google Scholar 

  • Smith, B. D., Perlstein, W. M., Davidson, R. A., & Michael, K. (1986). Sensation seeking: Differential effects of relevant, novel stimulation on electrodermal activity. Personality and Individual Differences, 7, 445–452.

    Article  Google Scholar 

  • Smith, B. D., Wilson, R. J., & Jones, B. E. (1983). Extraversion and multiple levels of caffeine-induced arousal: Effects on overhabituation and dishabituation. Psychophysiology, 20, 29–34.

    Article  PubMed  Google Scholar 

  • Sokolov, E. N. (1960). Neuronal models in the orienting reflex. In M. A. Brazier (Ed.), The central nervous system and behavior (pp. 187–275). New York: Macy Foundation.

    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 

  • Sokolov, E. N. (1966). Orienting reflex as information regulator. In A. N. Leontiev, A. R. Luria, E. N. Sokolov, & O. S. Vinogradova (Eds.), Psychological research in the USSR (Vol. 1, pp. 334–359). Moscow: Progress.

    Google Scholar 

  • Sokolov, E. N., & Boucsein, W. (2000). A psychophysiological model of emotion space. Integrative Physiological and Behavioral Science, 35, 81–119.

    Article  PubMed  Google Scholar 

  • Solanto, M. V., & Katkin, E. S. (1979). Classical EDR conditioning using a truly random control and subjects differing in electrodermal lability level. Bulletin of the Psychonomic Society, 14, 49–52.

    Google Scholar 

  • Solomon, K., & Hart, R. (1978). Pitfalls and prospects in clinical research on antianxiety drugs: Benzodiazepines and placebo: A research review. The Journal of Clinical Psychiatry, 39, 823–831.

    PubMed  Google Scholar 

  • Sorgatz, H. (1978). Components of skin impedance level. Biological Psychology, 6, 121–125.

    Article  PubMed  Google Scholar 

  • Sorgatz, H., & Pufe, P. (1978). Die differentielle Reagibilität der elektrodermalen Aktivität für aversive Reize. Zeitschrift für Experimentelle und Angewandte Psychologie, 3, 465–473.

    Google Scholar 

  • Sosnowski, T. (1988). Patterns of skin conductance and heart rate changes under anticipatory stress conditions. Journal of Psychophysiology, 2, 231–238.

    Google Scholar 

  • Sosnowski, T., Nurzynska, M., & Polec, M. (1991). Active-passive coping and skin conductance and heart rate changes. Psychophysiology, 28, 665–672.

    Article  PubMed  Google Scholar 

  • Sostek, A. J. (1978). Effects of electrodermal lability and payoff instructions on vigilance performance. Psychophysiology, 15, 561–568.

    Article  Google Scholar 

  • Spence, K. W., Haggard, D. F., & Ross, L. E. (1958). UCS intensity and the associative (habit) strength of the eyelid CR. Journal of Experimental Psychology, 55, 404–411.

    Article  PubMed  Google Scholar 

  • Spielberger, C. D. (1983). Manual for the State-Trait Anxiety Inventory. Palo Alto, CA: Consulting Psychologist.

    Google Scholar 

  • Spinks, J. A. (1977). Information and the orientating response. Unpublished Doctoral Dissertation, University of Southampton, Southampton

    Google Scholar 

  • Spinks, J. A., Blowers, G. H., & Shek, D. T. L. (1985). The role of the orienting response in the anticipation of information: A skin conductance response study. Psychophysiology, 22, 385–394.

    Article  PubMed  Google Scholar 

  • Spinks, J. A., & Siddle, D. (1983). The functional significance of the orienting response. In D. Siddle (Ed.), Orienting and habituation: Perspectives in human research (pp. 237–314). Chichester: Wiley.

    Google Scholar 

  • Spohn, H. E., & Patterson, T. (1979). Recent studies of psychophysiology in schizophrenia. Schizophrenia Bulletin, 5, 581–611.

    PubMed  Google Scholar 

  • Spohn, H. E., Thetford, P. E., & Cancro, R. (1971). The effects of phenothiazine medication on skin conductance and heart rate in schizophrenic patients. The Journal of Nervous and Mental Disease, 152, 129–139.

    Article  PubMed  Google Scholar 

  • Sponheim, S. R., Allen, J. J., & Iacono, W. G. (1995). Selected psychophysiological measures in depression: The significance of electrodermal Activity, electroencephalographic asymmetries, and contingent negative variation to behavioral and neurobiological aspects of depression. In G. A. Miller (Ed.), The behavioral high risk paradigm in psychopathology (pp. 222–249). New York: Springer.

    Chapter  Google Scholar 

  • Springer, S. P. (1977). Tachistoscopic and dichotic listening investigations of laterality in normal human subjects. In S. Harnad, R. W. Doty, L. Goldstein, J. Jaynes, & G. Krauthamer (Eds.), Lateralization in the nervous system (pp. 325–336). New York: Academic.

    Google Scholar 

  • Springer, S. P., & Deutsch, G. (1981). Left brain, right brain. San Francisco: Freeman.

    Google Scholar 

  • Springer, J., Müller, T., Langner, T., Luczak, H., & Beitz, W. (1990). Stress and strain caused by CAD-work – Results of a laboratory study. In L. Berlinguet & D. Berthelette (Eds.), Work with Display Units 1989 (pp. 231–238). Amsterdam: Elsevier.

    Google Scholar 

  • Squires, R. F., & Braestrup, C. (1977). Benzodiazepine receptors in rat brain. Nature, 266, 732–734.

    Article  PubMed  Google Scholar 

  • Steinberg, E. P., & Schwartz, G. E. (1976). Biofeedback and electrodermal self-regulation in psychopathy. Journal of Abnormal Psychology, 85, 408–415.

    Article  PubMed  Google Scholar 

  • Stellar, J. R., & Stellar, E. (1985). The neurobiology of motivation and reward. New York: Springer.

    Google Scholar 

  • Steller, M. (1987). Psychophysiologische Aussagebeurteilung. Wissenschaftliche Grundlagen und Anwendungsmöglichkeiten der “Lügendetektion.”. Göttingen: Hogrefe.

    Google Scholar 

  • Stelmack, R. M. (1981). The psychophysiology of extraversion and neuroticism. In H. J. Eysenck (Ed.), A model for personality (pp. 38–64). New York: Springer.

    Chapter  Google Scholar 

  • Stelmack, R. M., Plouffe, L., & Falkenberg, W. (1983). Extraversion, sensation seeking and electrodermal response: Probing a paradox. Personality and Individual Differences, 4, 607–614.

    Article  Google Scholar 

  • Stelmack, R. M., Plouffe, L. M., & Winogron, H. W. (1983). Recognition memory and the orienting response: An analysis of the encoding of pictures and words. Biological Psychology, 16, 49–63.

    Article  PubMed  Google Scholar 

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

    Google Scholar 

  • Stemmler, G. (1989). The autonomic differentiation of emotions revisited: Convergent and discriminant validation. Psychophysiology, 26, 617–631.

    Article  PubMed  Google Scholar 

  • Stemmler, G. (2002). Methodological considerations in the psychophysiological study of emotion. In R. J. Davidson, H. H. Goldsmith, & K. R. Scherer (Eds.), Handbook of affective science (pp. 225–255). New York: Oxford University Press.

    Google Scholar 

  • Stephenson, D., & Siddle, D. A. T. (1976). Effects of “below-zero” habituation on the electrodermal orienting response to a test stimulus. Psychophysiology, 13, 10–15.

    Article  PubMed  Google Scholar 

  • Stephenson, D., & Siddle, D. A. T. (1983). Theories of habituation. In D. Siddle (Ed.), Orienting and habituation: Perspectives in human research (pp. 183–236). Chichester: Wiley.

    Google Scholar 

  • Steptoe, A., & Greer, K. (1980). Relaxation and skin conductance feedback in the control of reactions to cognitive tasks. Biological Psychology, 10, 127–138.

    Article  PubMed  Google Scholar 

  • Steptoe, A., Melville, D., & Ross, A. (1984). Behavioral response demands, cardiovascular reactivity, and essential hypertension. Pyschosomatic Medicine, 46, 33–48.

    Google Scholar 

  • Steptoe, A., & Ross, A. (1981). Psychophysiological reactivity and the prediction of cardiovascular disorders. Journal of Psychosomatic Research, 25, 23–31.

    Article  PubMed  Google Scholar 

  • Stern, R. M. (1972). Detection of one’s own spontaneous GSRs. Psychonomic Science, 29, 354–356.

    Google Scholar 

  • Stern, J. A., & Janes, C. L. (1973). Personality and psychopathology. In W. F. Prokasy & D. C. Raskin (Eds.), Electrodermal activity in psychological research (pp. 283–346). New York: Academic.

    Google Scholar 

  • Stern, J. A., Surphlis, W., & Koff, E. (1965). Electrodermal responsiveness as related to psychiatric diagnosis and prognosis. Psychophysiology, 2, 51–61.

    Article  PubMed  Google Scholar 

  • Stevens, P. (2000). Human electrodermal response to remote human monitoring: Classification and analysis of response characteristics. The Journal of Parapsychology, 64, 391–409.

    Google Scholar 

  • Stewart, M. A., Stern, J. A., Winokur, G., & Fredman, S. (1961). An analysis of GSR conditioning. Psychological Review, 68, 60–67.

    Article  Google Scholar 

  • Stocksmeier, U., & Langosch, W. (1973). Die Galvanische Hautreaktion bei inneren Erkrankungen. Göttingen: Hogrefe.

    Google Scholar 

  • Stone, L. A., & Nielson, K. A. (2001). Intact physiological response to arousal with impaired emotional recognition in alexithymia. Psychotherapy and Psychosomatics, 70, 92–102.

    Article  PubMed  Google Scholar 

  • Storm, H., Myre, K., Rostrup, M., Stokland, O., Lien, M. D., & Raeder, J. C. (2002). Skin conductance correlates with perioperative stress. Acta Anesthesiolologica Scandinavica, 46, 887–895.

    Article  Google Scholar 

  • Storrie, M. C., Doerr, H. O., & Johnson, M. H. (1981). Skin conductance characteristics of depressed subjects before and after therapeutic intervention. The Journal of Nervous and Mental Disease, 169, 176–179.

    Article  PubMed  Google Scholar 

  • Straube, E. R. (1979). On the meaning of electrodermal nonresponding in schizophrenia. The Journal of Nervous and Mental Disease, 167, 601–611.

    Article  PubMed  Google Scholar 

  • Strelau, J. (1983). Temperament, personality, activity (pp. 141–166). New York: Academic.

    Google Scholar 

  • Sturgeon, D., Kuipers, L., Berkowitz, R., Turpin, G., & Leff, J. (1981). Psychophysiological responses of schizophrenic patients to high and low expressed emotion relatives. The British Journal of Psychiatry, 138, 40–45.

    Article  PubMed  Google Scholar 

  • Sturgeon, D., Turpin, G., Kuipers, L., Berkowitz, R., & Leff, J. (1984). Psychophysiological responses of schizophrenic patients to high and low expressed emotion relatives: A follow-up study. The British Journal of Psychiatry, 145, 62–69.

    Article  PubMed  Google Scholar 

  • Surwit, R. S., & Poser, E. G. (1974). Latent inhibition in the conditioned electrodermal response. Journal of Comparative and Physiological Psychology, 86, 543–548.

    Article  PubMed  Google Scholar 

  • Sutton, S. K., & Davidson, R. J. (1997). Prefrontal brain asymmetry: A biological substrate of the behavioral approach and inhibition systems. Psychological Science, 8, 204–210.

    Article  Google Scholar 

  • Suzuki, A., Hirota, A., Takasawa, N., & Shigemasu, K. (2003). Application of the somatic marker hypothesis to individual differences in decision-making. Biological Psychology, 65, 81–88.

    Article  PubMed  Google Scholar 

  • Tabbert, K., Stark, R., Kirsch, P., & Vaitl, D. (2006). Dissociation of neural responses and skin conductance reactions during fear conditioning with and without awareness of stimulus contingencies. NeuroImage, 32, 761–770.

    Article  PubMed  Google Scholar 

  • Takagi, K., & Nakayama, T. (1959). Peripheral effector mechanism of galvanic skin reflex. The Japanese Journal of Physiology, 9, 1–7.

    Article  PubMed  Google Scholar 

  • Tanaka, J., Ishida, S., Kawagoe, H., & Kondo, S. (2000). Workload of using a driver assistance system. IEEE Intelligent Transportation Systems Conference Proceedings .

    Google Scholar 

  • Tarrier, N., Cooke, E. C., & Lader, M. H. (1978). Electrodermal and heart-rate measurements in chronic and partially remitted schizophrenic patients. Acta Psychiatrica Scandinavica, 57, 369–376.

    Article  PubMed  Google Scholar 

  • Tarrier, N., Vaughn, C., Lader, M. H., & Leff, J. P. (1979). Bodily reactions to people and events in schizophrenics. Archives of General Psychiatry, 36, 311–315.

    PubMed  Google Scholar 

  • Taylor, J. A. (1953). A personality scale of manifest anxiety. Journal of Abnormal and Social Psychology, 48, 285–290.

    Article  Google Scholar 

  • Taylor, D. H. (1964). Drivers galvanic skin response and the risk of accident. Ergonomics, 7, 439–451.

    Article  Google Scholar 

  • Taylor, J., Carlson, S. R., Iacono, W. G., Lykken, D. T., & McGue, M. (1999). Individual differences in electrodermal responsivity to predictable aversive stimuli and substance dependence. Psychophysiology, 36, 193–198.

    Article  PubMed  Google Scholar 

  • Tchanturia, K., Liao, P. C., Uher, R., Lawrence, N., Treasure, J., & Campbell, I. C. (2007). An investigation of decision making in anorexia nervosa using the Iowa Gambling Task and skin conductance measurements. Journal of the International Neuropsychological Society, 13, 635–641.

    Article  PubMed  Google Scholar 

  • Thackray, R. I., & Orne, M. T. (1968). A comparison of physiological indices in detection of deception. Psychophysiology, 4, 329–339.

    Article  PubMed  Google Scholar 

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

    Google Scholar 

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

    Google Scholar 

  • Thom, E. (1988). Die Hamburger EDA-Auswertung. In W. Boucsein (Ed.), Elektrodermale Aktiviät. Grundlagen, Methoden und Anwendungen (pp. 501–514). Berlin: Springer.

    Google Scholar 

  • Thompson, J. K., & Adams, H. E. (1984). Psychophysiological characteristics of headache patients. Pain, 18, 41–52.

    Article  PubMed  Google Scholar 

  • Thompson, R. F., Groves, P. M., Teyler, T. J., & Roemer, R. A. (1973). A dual-process theory of habituation: Theory and behavior. In H. V. S. Peeke & M. J. Herz (Eds.), habituation (Behavioral studies, Vol. 1, pp. 239–271). New York: Academic.

    Google Scholar 

  • Thompson, R. F., & Spencer, W. A. (1966). Habituation: A model phenomenon for the study of neuronal substrates of behavior. Psychological Review, 73, 16–43.

    Article  PubMed  Google Scholar 

  • Thorell, L.-H., & d’Elia, G. (1988). Electrodermal activity in depressive patients in remission and in matched healthy subjects. Acta Psychiatrica Scandinavica 78 , 247–253.

    Google Scholar 

  • Thorpe, W. M. (1969). Learning and instinct in animals. London: Methuen.

    Google Scholar 

  • Thum, M., Boucsein, W. J., Kuhmann, W., & Ray, W. J. (1995). Standardized task strain and system response times in human-computer interaction. Ergonomics, 38, 1342–1351.

    Article  PubMed  Google Scholar 

  • Tomb, I., Hauser, M., Deldin, P., & Caramazza, A. (2002). Do somatic markers mediate decisions on the gambling task? Nature Neuroscience, 5, 1103–1104.

    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 

  • Tranel, D. T. (1983). The effects of monetary incentive and frustrative nonreward on heart rate and electrodermal activity. Psychophysiology, 20, 652–657.

    Article  PubMed  Google Scholar 

  • Tranel, D., & Damasio, A. R. (1988). Non-conscious face recognition in patients with face agnosia. Behavioural Brain Research, 30, 235–249.

    Article  PubMed  Google Scholar 

  • Tranel, D., & Damasio, A. R. (1993). The covert learning of affective valence does not require structures in hippocampal system or amygdala. Journal of Cognitive Neuroscience, 5, 79–88.

    Article  Google Scholar 

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

    Article  PubMed  Google Scholar 

  • Tranel, D. T., Fisher, A. E., & Fowles, D. C. (1982). Magnitude of incentive effects on heart rate. Psychophysiology, 19, 514–519.

    Article  PubMed  Google Scholar 

  • Tranel, D., Fowles, D. C., & Damasio, A. R. (1985). Electrodermal discrimination of familiar and unfamiliar faces: A methodology. Psychophysiology, 22, 403–408.

    Article  PubMed  Google Scholar 

  • Trasler, G. (1973). Criminal behaviour. In H. J. Eysenck (Ed.), Handbook of abnormal psychology (pp. 67–92). London: Pitman.

    Google Scholar 

  • Traxel, W. (1960). Die Möglichkeit einer objektiven Messung der Stärke von Gefühlen. Psychologische Forschung, 26, 75–90.

    PubMed  Google Scholar 

  • Tucker, D. M. (1981). Lateral brain function, emotion and conceptualization. Psychological Bulletin, 89, 19–46.

    Article  PubMed  Google Scholar 

  • Turkstra, L. S. (1995). Electrodermal response and outcome from severe brain injury. Brain Injury, 9, 61–80.

    Article  PubMed  Google Scholar 

  • Turpin, G. (1986). Effects of stimulus intensity on autonomic responding: The problem of differentiating orienting and defense reflexes. Psychophysiology, 23, 1–14.

    Article  PubMed  Google Scholar 

  • Turpin, G., Schaefer, F., & Boucsein, W. (1999). Effects of stimulus intensity, risetime, and duration on autonomic and behavioral responding: Implications for the differentiation of orienting, startle, and defense responses. Psychophysiology, 36, 453–463.

    Article  PubMed  Google Scholar 

  • Turpin, G., & Siddle, D. A. T. (1979). Effects of stimulus intensity on electrodermal activity. Psychophysiology, 16, 582–591.

    Article  PubMed  Google Scholar 

  • Turpin, G., Tarrier, N., & Sturgeon, D. (1988). Social psychophysiology and the study of biopsychosocial models of schizophrenia. In H. L. Wagner (Ed.), Social psychophysiology and emotion: Theory and clinical applications (pp. 251–272). Chichester: Wiley.

    Google Scholar 

  • Uchiyama, I. (1992). Differentiation of fear, anger, and joy. Perceptual and Motor Skills, 74, 663–667.

    PubMed  Google Scholar 

  • Undeutsch, U. (1983). Die psychophysiologische Täterschaftsermittlung. In F. Lösel (Ed.), Kriminalpsychologie (pp. 191–206). Weinheim: Beltz.

    Google Scholar 

  • Uno, T., & Grings, W. W. (1965). Autonomic components of orienting behavior. Psychophysiology, 1, 311–321.

    Article  PubMed  Google Scholar 

  • Van Bokhoven, I., Matthys, W., van Goozen, S. H. M., & van Engeland, H. (2005). Prediction of adolescent outcome in children with disruptive behaviour disorders. A study of neurobiological, psychological and family factors. European Child and Adolescent Psychiatry, 14, 153–163.

    Article  PubMed  Google Scholar 

  • Van Doornen, L. J. P., Orlebeke, J. F., & Somsen, J. M. (1980). Coronary risk and coping with aversive stimuli. Psychophysiology, 17, 598–603.

    Article  PubMed  Google Scholar 

  • Van Dyke, J. L., Rosenthal, D., & Rasmussen, P. V. (1974). Electrodermal functioning in adopted-away offspring of schizophrenics. Journal of Psychiatric Research, 10, 199–215.

    Article  PubMed  Google Scholar 

  • Van Goozen, S. H. M., Matthys, W., Cohen-Kettenis, P. T., Buitelaar, J. K., & van Engeland, H. (2000). Hypothalamic-pituitary-adrenal axis and autonomic nervous system activity in disruptive children and matched controls. Journal of the American Academy of Child and Adolescent Psychiatry, 39, 1438–1445.

    Article  PubMed  Google Scholar 

  • Vansteenwegen, D., Iberico, C., Vervliet, B., Marescau, V., & Hermans, D. (2008). Contextual fear induced by unpredictability in a human fear conditioning preparation is related to the chronic expectation of a threatening US. Biological Psychology, 77, 39–46.

    Article  PubMed  Google Scholar 

  • Vaughan, K. B., & Lanzetta, J. T. (1981). The effect of modification of expressive displays on skin reflex. Psychophysiology, 4, 223–228.

    Google Scholar 

  • Venables, P. H. (1975). Psychophysiological studies of schizophrenic pathology. In P. H. Venables & M. J. Christie (Eds.), Research in psychophysiology (pp. 282–324). London: Wiley.

    Google Scholar 

  • Venables, P. H. (1977). The electrodermal psychophysiology of schizophrenics and children at risk for schizophrenia: Controversies and developments. Schizophrenia Bulletin, 3, 28–48.

    PubMed  Google Scholar 

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

    Article  PubMed  Google Scholar 

  • Venables, P. H. (1983). Some problems and controversies in the psychophysiological investigation of schizophrenia. In A. Gale & J. A. Edwards (Eds.), Physiological correlates of human behaviour (Individual differences and psychopathology, Vol. 3, pp. 207–232). London: Academic.

    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 

  • Verschuere, B., Crombez, G., De Clercq, A., & Koster, E. H. W. (2004). Autonomic and behavioral responding to concealed information: Differentiating orienting and defensive responses. Psychophysiology, 41, 461–466.

    Article  PubMed  Google Scholar 

  • Vervliet, B., Vansteenwegen, D., Baeyens, F., Hermans, D., & Eelen, P. (2005). Return of fear in a human differential conditioning paradigm caused by a stimulus change after extinction. Behavior Research and Therapy, 43, 357–371.

    Article  Google Scholar 

  • Vetrugno, R., Liguori, R., Cortelli, P., & Montagna, P. (2003). Sympathetic skin response. Basic mechanisms and clinical applications. Clinical Autonomic Research, 13, 256–270.

    Article  PubMed  Google Scholar 

  • Vickland, V., Rogers, C., Craig, A., & Tran, Y. (2008). Electrodermal activity as a possible physiological marker for acupuncture. Complementary Therapies in Clinical Practice, 14, 83–89.

    Article  PubMed  Google Scholar 

  • Vögele, C., & Florin, I. (1997). Psychophysiological responses to food exposure: An experimental study in binge eaters. International Journal of Eating Disorders, 21, 147–157.

    Article  PubMed  Google Scholar 

  • Vögele, C., & Steptoe, A. (1986). Physiological and subjective stress reponses in surgical patients. Journal of Psychosomatic Research, 30, 205–215.

    Article  PubMed  Google Scholar 

  • Volow, M. R., Erwin, C. W., & Cipolat, A. L. (1979). Biofeedback control of skin potential level. Biofeedback and Self-Regulation, 4, 133–143.

    Article  PubMed  Google Scholar 

  • von Dultzig, K. (1997). Erfolgskontrolle von Fernsehwerbespots. Köln: Josef Eul.

    Google Scholar 

  • Vossel, G., & Roßmann, R. (1982). Interindividuelle Unterschiede in der Habituationsgeschwindigkeit der EDA: Eine systematische Analyse der Zusammenhänge verschiedener Habituationskennwerte. Zeitschrift für Differentielle und Diagnostische Psychologie, 3, 281–292.

    Google Scholar 

  • Vossel, G., & Roßmann, R. (1984). Electrodermal habituation speed and visual monitor performance. Psychophysiology, 21, 97–100.

    Article  PubMed  Google Scholar 

  • Wagner, A. R. (1969). Stimulus selection and a “modified continuity theory. In G. H. Bowers & J. T. Spence (Eds.), The psychology of learning and motivation (Vol. 3, pp. 1–41). New York: Academic.

    Google Scholar 

  • Wagner, A. R. (1976). Priming in STM: An information-processing mechanism for self-generated or retrieval-generated depression in performance. In T. J. Tighe & R. N. Leaton (Eds.), Habituation: Perspectives from child development, animal behavior, and neurophysiology (pp. 95–128). Hillsdale, NJ: Erlbaum.

    Google Scholar 

  • Wagner, A. R. (1978). Expectancies and the priming of STM. In S. H. Hulse, H. Fowler, & W. K. Honig (Eds.), Cognitive process in animal behavior (pp. 177–209). Hillsdale, NJ: Erlbaum.

    Google Scholar 

  • Wagner, H. (1989). The peripheral physiological differentiation of emotions. In H. Wagner & A. Manstead (Eds.), Handbook of social psychophysiology (pp. 77–98). London: Wiley.

    Google Scholar 

  • Waid, W. M., & Orne, M. T. (1980). Individual differences in electrodermal lability and the detection of information and deception. The Journal of Applied Psychology, 65, 1–8.

    Article  PubMed  Google Scholar 

  • Waid, W. M., & Orne, M. T. (1981). Cognitive, social, and personality processes in the physiological detection of deception. In L. Berkowitz (Ed.), Advances in experimental social psychology (pp. 61–106). New York: Academic.

    Google Scholar 

  • Waid, W. M., Orne, E. C., Cook, M. R., & Orne, M. T. (1981). Meprobamate reduces accuracy of physiological detection of deception. Science, 212, 71–73.

    Article  PubMed  Google Scholar 

  • Waid, W. M., Wilson, S. K., & Orne, M. T. (1981). Cross-modal physiological effects of electrodermal lability in the detection of deception. Journal of Personality and Social Psychology, 40, 1118–1125.

    Article  PubMed  Google Scholar 

  • Walrath, L. C., & Stern, J. A. (1980). General considerations. In H. M. van Praag, M. H. Lader, O. J. Rafaelsen, & E. J. Sachor (Eds.), Handbook of biological psychiatry: Part 2. Brain mechanisms and abnormal behavior – Psychophysiology (pp. 1–43). New York: Dekker.

    Google Scholar 

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

    Google Scholar 

  • Walschburger, P. (1986). Psychophysiological activation research. In J. Valsiner (Ed.), The individual subject and scientific psychology (pp. 311–345). London: Plenum.

    Google Scholar 

  • Wang, S. J., Fuh, J. L., Shan, D. E., Liao, K. K., Lin, K. P., Tsai, C. P., et al. (1993). Sympathetic skin response and R-R interval variation in Parkinson’s disease. Movement Disorders, 8, 151–157.

    Article  PubMed  Google Scholar 

  • Wang, S.-M., & Kain, Z. N. (2001). Auricular acupuncture: A potential treatment for anxiety. Anasthesia & Analgesia, 92, 548–553.

    Article  Google Scholar 

  • Warburton, D. M. (1983). Exploration in the neurochemistry of behaviour. In G. Davey (Ed.), Animal models of human behaviour (pp. 339–353). London: Wiley.

    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 

  • Watt, N. F., Anthony, E. J., Wynne, L. C., & Rolf, J. E. (Eds.). (1984). Children at risk for schizophrenia: A longitudinal perspective. London: Cambridge University Press.

    Google Scholar 

  • Wegner, W. (1981). Täterschaftsermittlung durch Polygraphie. Köln: Carl Heymanns.

    Google Scholar 

  • Weike, A. I., Schupp, H. T., & Hamm, A. O. (2007). Fear acquisition requires awareness in trace but not delay conditioning. Psychophysiology, 44, 170–180.

    Article  PubMed  Google Scholar 

  • Weinberger, D. A., Schwartz, G. E., & Davidson, R. J. (1979). Low-anxious, high-anxious, and repressive coping styles: Psychometric patterns and behavioral and physiological responses to stress. Journal of Abnormal Psychology, 88, 369–380.

    Article  PubMed  Google Scholar 

  • Weinstein, J., Averill, J. R., Opton, E. M., & Lazarus, R. S. (1968). Defensive style and discrepancy between self-report and physiological indexes of stress. Journal of Personality and Social Psychology, 10, 406–413.

    Article  PubMed  Google Scholar 

  • Wendt, J., Lotze, M., Weike, A. I., Hosten, N., & Hamm, A. O. (2008). Brain activation and defensive response mobilization during sustained exposure to phobia-related and other affective pictures in spider phobia. Psychophysiology, 45, 205–215.

    Article  PubMed  Google Scholar 

  • Westbury, H. R., & Neumann, D. L. (2008). Empathy-related responses to moving film stimuli depicting human and non-human animal targets in negative circumstances. Biological Psychology, 78, 66–74.

    Article  Google Scholar 

  • Whang, M. C. (2008). The emotional computer adaptive to human emotion. In J. Westerink, M. Ouwerkerk, T. Overbeek, F. Pasveer, & B. de Ruyter (Eds.), Probing experience: From academic research to commercial propositions (pp. 209–219). Dordrecht: Springer.

    Google Scholar 

  • Whang, M. C., Lim, J. S., & Boucsein, W. (2003). Preparing computers for affective communications: A psychophysiological concept and preliminary results. Human Factors, 45, 623–634.

    Article  PubMed  Google Scholar 

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

    Article  PubMed  Google Scholar 

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

    Article  PubMed  Google Scholar 

  • Wilhelm, F. H., Pfaltz, M. C., Gross, J. J., Mauss, I. B., Kim, S. I., & Wiederhold, B. K. (2005). Mechanisms of virtual reality exposure therapy: The role of the behavioral activation and behavioral inhibition systems. Applied Psychophysiology and Biofeedback, 30, 271–284.

    Article  PubMed  Google Scholar 

  • Wilhelm, F. H., & Roth, W. T. (1997). Acute and delayed effects of Alprazolam on flight phobics during exposure. Behaviour Research and Therapy, 35, 831–841.

    Article  PubMed  Google Scholar 

  • Wilhelm, F. H., & Roth, W. T. (1998). Taking the laboratory to the skies: Ambulatory assessment of self-report, autonomic, and the respiratory responses in flying phobia. Psychophysiology, 35, 596–606.

    Article  PubMed  Google Scholar 

  • Wilken, J. A., Smith, B. D., Tola, K., & Mann, M. (2000). Trait anxiety and prior exposure to non-stressful stimuli: Effects on psychophysiological arousal and anxiety. International Journal of Psychophysiology, 37, 233–242.

    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, K. M., Iacono, W. G., & Remick, R. A. (1985). Electrodermal activity among subtypes of depression. Biological Psychiatry, 20, 158–162.

    Article  PubMed  Google Scholar 

  • Williams, W. C., Parsons, R. L., & Strayer, D. L. (1981). Classical discrimination conditioning using the solutions to verbal and spatial problems as CSs: Bilateral measures of electrodermal excitation and inhibition. Psychophysiology, 18, 148–149.

    Google Scholar 

  • Williamson, P. S., Fowles, D. C., & Weinberger, M. (1985). Electrodermal potential and conductance measurements clinically discriminate between cystic fibrosis and control patients. Pediatric Research, 19, 810–814.

    Article  PubMed  Google Scholar 

  • Williamson, M. J., Paul, S. M., & Skolnick, P. (1978). Labelling of benzodiazepine receptors in vivo. Nature, 275, 551–553.

    Article  PubMed  Google Scholar 

  • Wilson, G. R. (1985). A simple device for the objective evaluation of peripheral nerve injuries. The Journal of Hand Surgery, 10, 324–330.

    PubMed  Google Scholar 

  • Wilson, G. D. (1990). Personality, time of day and arousal. Personality and Individual Differences, 11, 153–168.

    Article  Google Scholar 

  • Wilson, G. F. (2002). An analysis of mental workload in pilots during flight using multiple psychophysiological measures. The International Journal of Aviation Psychology, 12, 3–18.

    Article  Google Scholar 

  • Wilson, J. W. D., & Dykman, R. A. (1960). Background autonomic activity in medical students. Journal of Comparative and Physiological Psychology, 53, 405–411.

    Article  Google Scholar 

  • Wilson, K. G., & Graham, R. S. (1989). Electrodermal lability and visual information processing. Psychophysiology, 26, 321–328.

    Article  PubMed  Google Scholar 

  • Winton, W. M., Putnam, L. E., & Krauss, R. M. (1984). Facial and autonomic manifestations of the dimensional structure of emotion. Journal of Experimental Social Psychology, 20, 195–216.

    Article  Google Scholar 

  • Woodrough, R. E., Canti, G., & Watson, B. W. (1975). Electrical potential difference between basal cell carcinoma, benign inflammatory lesions and normal tissue. The British Journal of Dermatology, 92, 1–7.

    Article  PubMed  Google Scholar 

  • Wright, N. A. (1983). The cell proliferation kinetics of the epidermis. In L. A. Goldsmith (Ed.), Biochemistry and physiology of the skin (Vol. 1, pp. 203–229). New York: Oxford University Press.

    Google Scholar 

  • Wright, N., & McGown, A. (2001). Vigilance on the civil flight deck: Incidence of sleepiness and sleep during long-haul flights and associated changes in physiological parameters. Ergonomics, 44, 82–106.

    Article  PubMed  Google Scholar 

  • Wundt, W. (1896). Grundriß der Psychologie. Leipzig: Engelmann.

    Google Scholar 

  • Wynn, J. K., Dawson, M. E., & Schell, A. M. (2000). Discrete and continuous prepulses have differential effects on startle prepulse inhibition and skin conductance orienting. Psychophysiology, 37, 224–230.

    Article  PubMed  Google Scholar 

  • Yagi, A. (2000). Engineering psychophysiology in Japan. In R. W. Backs & W. Boucsein (Eds.), Engineering psychophysiology: Issues and applications (pp. 361–368). Mahwah, NJ: Lawrence Erlbaum.

    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 

  • Yoshino, A., Kimura, Y., Yoshida, T., Takahashi, Y., & Nomura, S. (2005). Relationships between temperament dimensions in personality and unconscious emotional responses. Biological Psychiatry, 57, 1–6.

    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 

  • Zahn, T. P. (1978). Sensitivity of measurement and electrodermal “nonresponding” in schizophrenic and normal subjects. Schizophrenia Bulletin, 4, 153.

    PubMed  Google Scholar 

  • Zahn, T. P. (1986). Psychophysiological approaches to psychopathology. In M. G. H. Coles, E. Donchin, & S. W. Porges (Eds.), Psychophysiology: Systems, processes, applications (pp. 508–610). Amsterdam: Elsevier.

    Google Scholar 

  • Zahn, T. P., Carpenter, W. T., & McGlashan, T. H. (1981a). Autonomic nervous system activity in acute schizophrenia: I. Method and comparison with normal controls. Archives of General Psychiatry, 38, 251–258.

    PubMed  Google Scholar 

  • Zahn, T. P., Carpenter, W. T., & McGlashan, T. H. (1981b). Autonomic nervous system activity in acute schizophrenia: II. Relationships to short-term prognosis and clinical state. Archives of General Psychiatry, 38, 260–266.

    PubMed  Google Scholar 

  • Zahn, T. P., Grafman, J., & Tranel, D. (1999). Frontal lobe lesions and electrodermal activity: Effects of significance. Neuropsychologia, 37, 1227–1241.

    Article  PubMed  Google Scholar 

  • Zahn, T. P., Rosenthal, D., & Lawlor, W. G. (1968). Electrodermal and heart rate orienting reactions in chronic schizophrenia. Journal of Psychiatric Research, 6, 117–134.

    Article  PubMed  Google Scholar 

  • Zeier, H. (1979). Concurrent physiological activity of driver and passenger when driving with and without automatic transmission in heavy city traffic. Ergonomics, 22, 799–810.

    Article  PubMed  Google Scholar 

  • Zeiner, A. R. (1970). Orienting response and discrimination conditioning. Physiology and Behaviour, 5, 641–646.

    Article  Google Scholar 

  • Zoccolotti, P., Caltagirone, C., Pecchinenda, A., & Troisi, E. (1993). Electrodermal activity in patients with unilateral brain damage. In J.-C. Roy, W. Boucsein, C. D. Fowles, & J. H. Gruzelier (Eds.), Progress in Electrodermal Research (pp. 311–326). New York: Plenum.

    Chapter  Google Scholar 

  • Zoccolotti, P., Scabini, D., & Violani, C. (1982). Electrodermal responses in patients with unilateral brain damage. Journal of Clinical Neuropsychology, 4, 143–150.

    Article  PubMed  Google Scholar 

  • Zubin, J., & Spring, B. (1977). Vulnerability – A new view of schizophrenia. Journal of Abnormal Psychology, 86, 103–126.

    Article  PubMed  Google Scholar 

  • Zuckerman, M. (1983). A biological theory of sensation seeking. In M. Zuckerman (Ed.), Biological bases of sensation seeking, impulsivity, and anxiety (pp. 37–76). Hillsdale, NJ: Erlbaum.

    Google Scholar 

  • Zuckerman, M. (1990). The psychophysiology of sensation seeking. Journal of Personality, 58, 313–345.

    Article  PubMed  Google Scholar 

  • Zuckerman, M., Kolin, E. A., Price, L., & Zoob, I. (1964). Development of a sensation-seeking-scale. Journal of Consulting Psychology, 28, 477–482.

    Article  Google Scholar 

  • Zuckerman, M., & Lubin, B. (Eds.). (1965). Manual for the multiple affect adjective check list. San Diego: Edits.

    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). Applications of Electrodermal Recording. In: Electrodermal Activity. Springer, Boston, MA. https://doi.org/10.1007/978-1-4614-1126-0_3

Download citation

Publish with us

Policies and ethics