首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Emotional expressions influence social judgments of personality traits. The goal of the present research was to show that it is of interest to assess the impact of neutral expressions in this context. In 2 studies using different methodologies, the authors found that participants perceived men who expressed neutral and angry emotions as higher in dominance when compared with men expressing sadness or shame. Study 1 showed that this is also true for men expressing happiness. In contrast, women expressing either anger or happiness were perceived as higher in dominance than were women showing a neutral expression who were rated as less dominant. However, sadness expressions by both men and women clearly decreased the extent to which they were perceived as dominant, and a trend in this direction emerged for shame expressions by men in Study 2. Thus, neutral expressions seem to be perceived as a sign of dominance in men but not in women. The present findings extend our understanding of the way different emotional expressions affect perceived dominance and the signal function of neutral expressions—which in the past have often been ignored. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

2.
The authors conducted 2 studies to identify the vocal acoustical correlates of unresolved anger and sadness among women reporting unresolved anger toward an attachment figure. In Study 1, participants (N = 17) were induced to experience and express anger then sadness or sadness then anger. In Study 2, a 2nd group of participants (N = 22) underwent a relationship-oriented, emotion-focused analogue therapy session. Results from both studies showed that, relative to emotionally neutral speech, anger evoked an increase in articulation rate and in mean fundamental frequency (F0) and F0-range, whereas sadness evoked an increase in F0-perturbation. Both F0 and F0-range were larger for anger than for sadness. In addition, results from the mood-induction-procedure study revealed 2 Emotion×Order interactions. Whereas variations in amplitude range suggested that anger evoked less physiological activation when induced after sadness, variations in F0-perturbation suggested that sadness evoked more physiological activation when induced after anger. These findings illustrate the feasibility of using acoustical measures to identify clients' personally and clinically meaningful emotional experiences, and shifts between such emotional experiences, in the context of psychotherapy. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

3.
In this set of studies, we examine the perceptual similarities between emotions that share either a valence or a motivational direction. Determination is a positive approach-related emotion, whereas anger is a negative approach-related emotion. Thus, determination and anger share a motivational direction but are opposite in valence. An implemental mind-set has previously been shown to produce high-approach-motivated positive affect. Thus, in Study 1, participants were asked to freely report the strongest emotion they experienced during an implemental mind-set. The most common emotion reported was determination. On the basis of this result, we compared the facial expression of determination with that of anger. In Study 2, naive judges were asked to identify photographs of facial expressions intended to express determination, along with photographs intended to express basic emotions (joy, anger, sadness, fear, disgust, neutral). Correct identifications of intended determination expressions were correlated with misidentifications of the expressions as anger but not with misidentifications as any other emotion. This suggests that determination, a high-approach-motivated positive affect, is perceived as similar to anger. In Study 3, naive judges quantified the intensity of joy, anger, and determination expressed in photographs. The intensity of perceived determination was directly correlated with the intensity of perceived anger (a high-approach-motivated negative affect) and was inversely correlated with the intensity of perceived joy (a low-approach-motivated positive affect). These results demonstrate perceptual similarity between emotions that share a motivational direction but differ in valence. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

4.
This study was designed to identify physiological correlates of unresolved anger and sadness, and the shift between these emotions, in a context similar to that of emotion-focused, experiential psychotherapy. Twenty-seven university students reporting unresolved anger toward an attachment figure were induced to experience and express unresolved anger and sadness. Simultaneously, their heart rate variability, finger temperature, and skin conductance levels were monitored. The sequence of emotion induction was counterbalanced. Sympathetic activation, as reflected by finger temperature, increased significantly from anger to sadness, but not from sadness to anger. A follow-up study (N=36) of participants induced to experience and express either anger or sadness in both the 1st and 2nd inductions ruled out an Anger×Time interaction and a sadness-sadness effect, suggesting that the increase in sympathetic activation from anger to sadness was a function of the unique sequence of emotions. These findings represent a first step toward using physiological measures to capture shifts from unresolved anger to vulnerable primary emotions during a therapy-like task and provide evidence for the purported mechanism underlying unresolved anger. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

5.
Do expressions of anger in conflict elicit competition or cooperation? To reconcile inconsistent results obtained in previous research, the authors developed and tested a dual-process model that proposes that power and the appropriateness of the expressions of anger jointly determine whether an individual facing an angry antagonist competes by demanding value or cooperates by conceding value. In a scenario study and a computer-mediated negotiation simulation, (a) participants with lower power claimed less value from an angry adversary than from a nonemotional one, regardless of the appropriateness of the expressions of anger, and (b) participants with higher power demanded more value when the adversary's expressions of anger were inappropriate than when they were appropriate or when the adversary was nonemotional. The theoretical and practical implications of the model and findings are discussed. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

6.
On separate visits to the laboratory, 36 nine-month-old infants (18 boys and 18 girls) watched their mothers express joy or sadness, facially and vocally, during a 2-min emotion-induction period. After the induction period, mothers continued to express joy or sadness while their infants played with four sets of toys. Infant emotion expressions were analyzed using the Max (Izard, 1979a) and Affex (Izard, Dougherty, & Hembree, 1983) coding systems, and infant play behavior was coded with a system developed by Belsky and Most (1981). The amount of time that the infants looked at their mothers was also measured. Findings were generally consistent with differential emotions theory (Izard, 1979b). The infants expressed more joy and looked longer at their mothers during the joy condition and they showed more sadness, anger, and gaze aversion during the sadness condition. The infants engaged in more play behavior in the joy condition than in the sadness condition. Regression analyses revealed several significant relations between infants' gaze behavior, emotion expressions, and play behavior. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

7.
This study investigated the identification of facial expressions of emotion in currently nondepressed participants who had a history of recurrent depressive episodes (recurrent major depression; RMD) and never-depressed control participants (CTL). Following a negative mood induction, participants were presented with faces whose expressions slowly changed from neutral to full intensity. Identification of facial expressions was measured by the intensity of the expression at which participants could accurately identify whether faces expressed happiness, sadness, or anger. There were no group differences in the identification of sad or angry expressions. Compared with CTL participants, however, RMD participants required significantly greater emotional intensity in the faces to correctly identify happy expressions. These results indicate that biases in the processing of emotional facial expressions are evident even after individuals have recovered from a depressive episode. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

8.
Sad and anxious feelings are known to increase in the immediate postpartum period, whereas studies on new mothers' other emotional qualities such as anger are scarce. In laboratory studies, attachment security was found to be associated with effective emotion regulation in challenging situations. This study investigated attachment representations of experiences with parents and of current experiences with the partner as predictors of sad, anxious, and angry feelings across the transition to motherhood. Seventy-seven pregnant women in their third trimester were administered the Adult Attachment Interview and the Current Relationship Interview. The Differential Emotions Scale was given in pregnancy and at the infant's ages of 2 weeks, 2, 4, and 6 months, asking both mothers and fathers about maternal emotional experience. Sadness and anxiety increased 2 weeks postpartum and returned to below baseline over the following months, while anger did not change. Contrary to mothers with an insecure representation of their couple relationship, those with a secure representation reported and displayed increased sadness and anxiety 2 weeks after giving birth, from which they quickly recovered. For mothers secure in their representation of past attachment relationships with parents, an increase of low-level anger emerged 4 months postpartum, which did not occur in insecure participants and receded quickly. It can be concluded that secure representations of current and past attachment relationships help new mothers express and recover from negative emotions. These findings further elucidate the associations between attachment status and emotion regulation while adding a couple perspective. (PsycINFO Database Record (c) 2011 APA, all rights reserved)  相似文献   

9.
This study examined the influence of socialization figures (mother, father, best friend, medium friend), emotion type (anger, sadness, physical pain), age, and gender on 66 2nd and 71 5th-grade children's reasons for and methods of affect expression. Children reported expressing sadness in order to receive support, expressing pain because they perceived it was uncontrollable, and regulating anger due to negative consequences. Girls reported using verbal means to communicate emotion, whereas boys cited mild aggressive methods. Younger children indicated expressing emotion to receive assistance because they lack regulation skills, and to adhere to norms. Children expressed emotion in passive ways to fathers more than peers, and mothers were deemed by younger children as most accepting of displays of anger. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

10.
The authors compared the accuracy of emotion decoding for nonlinguistic affect vocalizations, speech-embedded vocal prosody, and facial cues representing 9 different emotions. Participants (N = 121) decoded 80 stimuli from 1 of the 3 channels. Accuracy scores for nonlinguistic affect vocalizations and facial expressions were generally equivalent, and both were higher than scores for speech-embedded prosody. In particular, affect vocalizations showed superior decoding over the speech stimuli for anger, contempt, disgust, fear, joy, and sadness. Further, specific emotions that were decoded relatively poorly through speech-embedded prosody were more accurately identified through affect vocalizations, suggesting that emotions that are difficult to communicate in running speech can still be expressed vocally through other means. Affect vocalizations also showed superior decoding over faces for anger, contempt, disgust, fear, sadness, and surprise. Facial expressions showed superior decoding scores over both types of vocal stimuli for joy, pride, embarrassment, and “neutral” portrayals. Results are discussed in terms of the social functions served by various forms of nonverbal emotion cues and the communicative advantages of expressing emotions through particular channels. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

11.
The term “humiliated fury” refers to the anger people can experience when they are shamed. In Study 1, participants were randomly exposed to a prototypical shameful event or control event, and their self-reported feelings of anger were measured. In Study 2, participants reported each school day, for 2 weeks, the shameful events they experienced. They also nominated classmates who got angry each day. Narcissism was treated as a potential moderator in both studies. As predicted, shameful events made children angry, especially more narcissistic children. Boys with high narcissism scores were especially likely to express their anger after being shamed. These results corroborate clinical theory holding that shameful events can initiate instances of humiliated fury. (PsycINFO Database Record (c) 2011 APA, all rights reserved)  相似文献   

12.
Persistence of instrumental responding and negative facial expressions in response to repeated goal blockage was studied in 53 4-month-old infants. All participants experienced 2 sessions comprising baseline (no stimulation), contingency (stimulation resulting from infant action), and extinction (no stimulation) on consecutive days. Performance criteria identified 2 groups of infants, those who learned in Session 1 (Learning Group 1) and those who learned in Session 2 (Learning Group 2). Individual differences in instrumental responses and facial expression during extinction were compared as a function of learning group. Across sessions, the repetition of extinction for Learning Group 1 was associated with both a persistent instrumental response and anger expressions. The level of instrumental response and anger expression was equivalent to that observed for Learning Group 2 but only in Session 2, the day on which that group learned. Sadness and anger/sadness blended expressions were initially more common in Learning Group 2, but these expressions were attenuated given another exposure to the contingency in Session 2. Implications for the relations among infant emotion, cognition, and behavior are discussed. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

13.
The facial expressions of 28 13-mo-old middle-class children were videotaped during the 3-min separation episode of the Ainsworth strange-situation procedure (ASSP). Facial behavior was analyzed to determine the patterns of emotional expressions during separation and to assess the relations between these patterns and types of attachment as assessed by the ASSP. Findings reveal that anger was the dominant negative emotion expressed by the majority of Ss in each of 3 ad hoc groups determined by level of negative emotion. Some high-negative emotion expressers displayed predominantly anger and others mainly sadness. Patterns of emotion expression varied with type of attachment; Ss who showed an insecure-resistant attachment pattern displayed less interest and more sadness than Ss in the securely attached groups. The proportion of time anger was expressed did not differ significantly with type of attachment. (20 ref) (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

14.
Parents were asked to recall recent events that had evoked happiness, sadness, anger, and fear in their children. Children (N?=?77, 2 years 3 months to 6 years 6 months) indicated whether they remembered each event, and if so, they described the event and how it had made them feel. Agreement between parent and child concerning how the child felt varied as a function of emotion. Children agreed with their parents' emotion attributions most often for events that parents recalled as having evoked happiness and sadness, less often for fear, and least often for anger. Children disagreed with parents' attributions of happiness and sadness most often when parents and children differed concerning the attribution of children's goals. Discordant reports about children's anger were most frequent when parents and children reported conflicting goals. Discordant reports about fear were most frequent when parents and children focused on different parts of the temporal sequence surrounding the event. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

15.
Scholars have argued that anger expressed by participants in mediation is counterproductive; yet, there is also reason to believe that expressions of anger can be productive. The authors tested these competing theories of emotion by using data from online mediation. Results show that expression of anger lowers the resolution rate in mediation and that this effect occurs in part because expressing anger generates an angry response by the other party. However, when respondents are especially vulnerable, expressions of anger by the filer do not hinder settlement. The authors also examined precursors to anger, such as value of dispute and reputation, and the degree to which a focus on dispute resolution is reciprocated. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

16.
Emotion theorists assume certain facial displays to convey information about the expresser's emotional state. In contrast, behavioral ecologists assume them to indicate behavioral intentions or action requests. To test these contrasting positions, over 2,000 online participants were presented with facial expressions and asked what they revealed--feeling states, behavioral intentions, or action requests. The majority of the observers chose feeling states as the message of facial expressions of disgust, fear, sadness, happiness, and surprise, supporting the emotions view. Only the anger display tended to elicit more choices of behavioral intention or action request, partially supporting the behavioral ecology view. The results support the view that facial expressions communicate emotions, with emotions being multicomponential phenomena that comprise feelings, intentions, and wishes. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

17.
Two studies provided direct support for a recently proposed dialect theory of communicating emotion, positing that expressive displays show cultural variations similar to linguistic dialects, thereby decreasing accurate recognition by out-group members. In Study 1, 60 participants from Quebec and Gabon posed facial expressions. Dialects, in the form of activating different muscles for the same expressions, emerged most clearly for serenity, shame, and contempt and also for anger, sadness, surprise, and happiness, but not for fear, disgust, or embarrassment. In Study 2, Quebecois and Gabonese participants judged these stimuli and stimuli standardized to erase cultural dialects. As predicted, an in-group advantage emerged for nonstandardized expressions only and most strongly for expressions with greater regional dialects, according to Study 1. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

18.
This study compared young and older adults’ ability to recognize bodily and auditory expressions of emotion and to match bodily and facial expressions to vocal expressions. Using emotion discrimination and matching techniques, participants assessed emotion in voices (Experiment 1), point-light displays (Experiment 2), and still photos of bodies with faces digitally erased (Experiment 3). Older adults’ were worse at least some of the time in recognition of anger, sadness, fear, and happiness in bodily expressions and of anger in vocal expressions. Compared with young adults, older adults also found it more difficult to match auditory expressions to facial expressions (5 of 6 emotions) and bodily expressions (3 of 6 emotions). (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

19.
Across 6 studies, factors signaling potential vulnerability to harm produced a bias toward outgroup categorization—a tendency to categorize unfamiliar others as members of an outgroup rather than as members of one's ingroup. Studies 1 through 4 demonstrated that White participants were more likely to categorize targets as Black (as opposed to White) when those targets displayed cues heuristically associated with threat (masculinity, movement toward the perceiver, and facial expressions of anger). In Study 5, White participants who felt chronically vulnerable to interpersonal threats responded to a fear manipulation by categorizing threatening (angry) faces as Black rather than White. Study 6 extended these findings to a minimal group paradigm, in which participants who felt chronically vulnerable to interpersonal threats categorized threatening (masculine) targets as outgroup members. Together, findings indicate that ecologically relevant threat cues within both the target and the perceiver interact to bias the way people initially parse the social world into ingroup vs. outgroup. Findings support a threat-based framework for intergroup psychology. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

20.
Studied the effect of maternal facial expressions of emotion on 108 12-mo-old infants in 4 studies. The deep side of a visual cliff was adjusted to a height that produced no clear avoidance and much referencing of the mother. In Study 1, 19 Ss viewed a facial expression of joy, while 17 Ss viewed one of fear. In Study 2, 15 Ss viewed interest, while 18 Ss viewed anger. In Study 3, 19 Ss viewed sadness. In Study 4, 23 Ss were used to determine whether the expressions influenced Ss' evaluation of an ambiguous situation or whether they were effective in controlling behavior merely because of their discrepancy or unexpectedness. Results show that Ss used facial expressions to disambiguate situations. If a mother posed joy or interest while S referenced, most Ss crossed the deep side. If a mother posed fear or anger, few Ss crossed. In the absence of any depth whatsoever, few Ss referenced the mother and those who did, while the mother was posing fear, hesitated but crossed nonetheless. It is suggested that facial expressions regulate behavior most clearly in contexts of uncertainty. (17 ref) (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号