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 15 毫秒
1.
The purpose of the present study was to examine the associations of temperament and love attitudes with eating behaviors in 190 college-aged nonclinical women who completed a survey that included measures of temperament, loving-style, and eating attitudes. Certain temperament and loving-style variables showed significant statistical association with scores on eating attitudes. Specifically, both obsessive and game-playing love-styles were related to the Dieting and Bulimia-Food Preoccupation dimensions of the eating scales, while temperamental fear and anger were related to bulimia and oral control. The role of interpersonal relationships and self-perceptions of temperament may provide a useful perspective for understanding the etiology of eating disorders.  相似文献   

2.
We examined the association between adulthood emotionality-activity-sociability temperament scale and preclinical atherosclerosis and, whether this association is mediated by cardiovascular risk factors (low-density lipoprotein cholesterol, systolic blood pressure, diastolic blood pressure and body-mass index (BMI)). The participants were a nationally representative sample of 537 men and 811 women from the Cardiovascular Risk in Young Finns study aged 15-30 years at the baseline in 1992 and aged 24-39 years at the follow-up in 2001. Carotid atherosclerosis was assessed by ultrasound scans of the common carotid artery intima-media thickness (IMT) and brachial flow-mediated dilation (FMD). In men, there was an association between the temperament dimension activity and IMT (β = 0.08, p = 0.036) which was partially mediated by BMI (β decreased from 0.08 to 0.05; p-value of Sobel test = 0.002). However, after correction for multiple comparisons the association between IMT and the temperament dimension activity in men was only of borderline significance. In women, there were no associations between temperament and IMT or FMD. These results suggest that a highly active temperament may contribute to early signs of atherosclerosis in men and that body mass may mediate this association.  相似文献   

3.
This article focuses on the relation between affect intensity and 3 fundamental dimensions of temperament—emotionality, sociability, and sensory arousability. The purpose was to show that individual differences in affect intensity as a dimension of temperament can influence not only advertising responses, but also the lifestyles and preferences of consumers. Study 1 confirmed the emotionality dimension in that high affect intensity individuals responded with significantly stronger levels of emotion when exposed to an affectively charged advertising appeal, but not when exposed to a nonemotional appeal. Studies 2 and 3 demonstrated that the fundamental dimensions of temperament are accompanied by heightened emotional intensity and do predict different preferences for lifestyle activities for high and low affect intensity consumers. A significant Affect Intensity × Gender interaction occurred indicating that both men and women expressed stronger emotions when experiencing activities that were gender‐congruent (e.g., watching sports on TV for men, and smelling perfumes for women). Future research directions are also discussed.  相似文献   

4.
This study examined associations of temperament at ages 6 to 12 with body-mass index (BMI) and waist circumference (WC) at ages 24 to 30 years. The participants were 619 men and women derived from the population-based Cardiovascular Risk in Young Finns Study. Temperament was operationalized as (negative) emotionality, sociability, and activity. High emotionality predicted increased BMI, independently of WC, and independently of childhood and adulthood risk factors for adult obesity. None of the temperament dimensions had any associations with WC after controlling for BMI. The findings suggest that temperamental difficulty in childhood may be a useful risk indicator for general body mass in adulthood, and the mechanisms relating temperament with body mass should be further explored.  相似文献   

5.
We explored how parent gender, infant temperament, and coparenting dynamics worked together to shape mothers’ and fathers’ depressive symptoms, stress, and parental efficacy during early parenthood. We were interested in the coparenting relationship as a context that shapes how parents respond to their infant's temperamental qualities. Participants were 139 couples who had recently given birth to their first child. Parent reports of temperament were collected when the infant was 4–8 months old and reports of coparenting and parent adjustment were collected at 13 months. Two-level random intercept models revealed interactions between temperament and coparenting, highlighting the family system as a context for how men and women adapt to their parenting role. There was little evidence for mother–father differences in these associations.  相似文献   

6.
Temperament and physical activity (PA) have been examined in children and adolescents, but little is known about these associations in adulthood. Personality traits, however, are known to contribute to PA in adults. This study, which examined both temperament and personality characteristics at age 42 in relation to frequency of PA at age 50 (JYLS, n = 214–261), also found associations with temperament traits. Positive associations were found between Orienting sensitivity and overall PA and between Extraversion and vigorous PA among women and between low Negative affectivity and overall and vigorous PA among men. Furthermore, Orienting sensitivity and Agreeableness were associated with vigorous PA among men. Temperament and personality characteristics also showed gender-specific associations with rambling in nature and watching sports.  相似文献   

7.
ABSTRACT We examined the relationship of Cloninger's temperament factors—Novelty Seeking, Harm Avoidance, Reward Dependence, and Persistence—to perceived threat and stress and performance appraisals during different challenges, i.e., mental arithmetic, the reaction time task, and three public speaking tasks, among 97 young adult men and women. Temperament was measured by the Temperament and Character Inventory. The results showed that, although some of the predictions made by Cloninger's model were confirmed, some were unsupported. The results revealed also some associations between temperament and cognitive appraisals that were intelligible, but not predicted by Cloninger's model. There were considerable domain specificity and gender differences in the associations found. Cloninger's temperament dimensions are related to threat, stress, and performance appraisals, thereby influencing individual's stress vulnerability, adjustment, and personal functioning.  相似文献   

8.
The present investigation tested the hypothesis that executive functioning (EF) would mediate the relation between difficult temperament (DT) and aggressive behavior. This model was tested in 310 adult men and women. DT was measured using the Dimensions of Temperament Scale—Revised, EF was measured using 7 well-established neuropsychological tests, and aggression was assessed using the Buss–Perry Aggression Questionnaire. EF successfully mediated the DT–aggression relation for men, however, the model did not hold for women. Results are discussed with regard to how they influence current models of aggressive behavior as well as their implications for future violence prevention efforts.  相似文献   

9.
ABSTRACT Earned Income and work accomplishment were determined at age 41 for 89 adults whose mothers had been interviewed for their child-rearing practices when the adults were 5 years old. At age 31, in spontaneous thought but not self-report, n Achievement predicted earned income and socialized power motivation predicted work accomplishment at age 41. Hardships (or "bad breaks") during childhood and adolescence predicted work outcomes for both men and women, as did education for men. Parenting achievement pressure in the first 2 years of life was associated with adult n Achievement and earned income, while moderate encouragement of assertiveness by mothers who were warm to boys and cool to girls was associated with adult socialized power motivation and work accomplishment. Controls for social class of origin, IQ, temperament, and education did not explain the relations between parenting, motivation, and work outcomes, although education played a larger role for men than women who worked both inside and outside the home.  相似文献   

10.
Several theories have been proposed to account for the apparent non-responsiveness to punishment cues or aversive events demonstrated by members of some disinhibited groups. Included among these theories are those that emphasize individual differences in temperament, temperament-related biases associated with the allocation of attentional resources, and impairments in executive functions. This study examined the relative contribution of each of these variables to the prediction of passive avoidance errors (PAEs, or failures to inhibit responding to punishment cues) during a computerized go/no-go task. Variations in temperament, attentional allocation to punishment feedback, and executive functions were found to independently and additively contribute to the prediction of PAEs in a mixed sample of men and women recruited at a university campus (n=145). Results from this study, therefore, support multiple theoretical perspectives on PAEs as assessed by the go/no-go experimental paradigm.  相似文献   

11.
Migration is a central determinant of population dynamics and structure. We examined whether three major temperament traits--sociability, emotionality, and activity--predicted migration propensity, selective urban-rural migration, and migration distance in a 9-year prospective study in Finland. The participants were Finnish women and men (N= 1,733) ages 15 to 30 years at baseline. The home municipality's position on the urban-rural continuum was assessed on the basis of the municipality's population density. We found that high sociability predicted migration to urban areas and longer migration distances. High activity increased general migration propensity (including migration to both urban and rural areas). High emotionality increased the likelihood of leaving the home municipality and decreased migration distances, but was not associated with selective urban-rural migration. These data suggest that temperament predicts the self-selection of environments on a demographic scale and may be relevant in understanding population dynamics.  相似文献   

12.
We used a prospective longitudinal design to examine the relation between parenting experiences at age 5 and level of self-criticism at age 12 and the stability of self-criticism from age 12 to age 31 in 156 subjects. The results showed that mothers' reports of parenting behaviors that reflect restrictiveness and rejection were related to the development of self-criticism, particularly when received from the same-sex parent. Partial correlational analyses revealed that the parenting-self-criticism relations remained significant when the mother's report of the child's early temperament was statistically controlled. The results also showed that for women, self-criticism was very stable from early adolescence to young adulthood. By contrast, there was no relation between self-criticism at ages 12 and 31 for men; however, there was a strong relation for men between age 12 self-criticism and inhibited aggressive impulses at age 31.  相似文献   

13.
Emerging evidence suggests that temperament may predict childbearing. We examined the association between four temperament traits (novelty seeking, harm avoidance, reward dependence and persistence of the Temperament and Character Inventory) and childbearing over the life course in the population‐based Cardiovascular Risk in Young Finns study (n = 1535; 985 women, 550 men). Temperament was assessed when the participants were aged 20–35 and fertility history from adolescence to adulthood was reported by the participants at age 30–45. Discrete‐time survival analysis modelling indicated that high childbearing probability was predicted by low novelty seeking (standardized OR = 0.92; 95% confidence interval 0.88–0.97), low harm avoidance (OR = 0.90; 0.85–0.95), high reward dependence (OR = 1.09; 1.03–1.15) and low persistence (OR = 0.91; 0.87–0.96) with no sex differences or quadratic effects. These associations grew stronger with increase in numbers of children. The findings were substantially the same in a completely prospective analysis. Adjusting for education did not influence the associations. Despite its negative association with overall childbearing, high novelty seeking increased the probability of having children in participants who were not living with a partner (OR = 1.29; 1.12–1.49). These data provide novel evidence for the role of temperament in influencing childbearing, and suggest possible weak natural selection of temperament traits in contemporary humans. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

14.
In this functional magnetic resonance imaging (fMRI) study we examined neural processing of infant faces associated with a happy or a sad temperament in nulliparous women. We experimentally manipulated adult perception of infant temperament in a probabilistic learning task. In this task, participants learned about an infant's temperament through repeated pairing of the infant face with positive or negative facial expressions and vocalizations. At the end of the task, participants were able to differentiate between “mostly sad” infants who cried often and “mostly happy” infants who laughed often. Afterwards, brain responses to neutral faces of infants with a happy or a sad temperament were measured with fMRI and compared to brain responses to neutral infants with no temperament association. Our findings show that a brief experimental manipulation of temperament can change brain responses to infant signals. We found increased amygdala connectivity with frontal regions and the visual cortex, including the occipital fusiform gyrus, during the perception of infants with a happy temperament. In addition, amygdala connectivity was positively related to the post-manipulation ratings of infant temperament, indicating that amygdala connectivity is involved in the encoding of the rewarding value of an infant with a happy temperament.  相似文献   

15.
Caring for infants with negative reactive temperament may tax parents' confidence in their caregiving ability, or parenting self‐efficacy (PSE). This may happen in particular in parents who interpret these signals as negative feedback on their performance. To test this hypothesis, 179 first‐time pregnant women were presented a caregiving simulation that provided positive and negative feedback on their attempts to comfort a crying baby. According to their PSE resilience to negative feedback during the task, they were grouped in a high resilient and low resilient group. PSE was followed up at 32 weeks of pregnancy and 3 and 12 months after birth, while perceived temperament of the child was assessed at 3 and 12 months after birth. Results showed that among women with low resilience against negative feedback, perceived negative temperament was negatively associated with PSE at 3 months, whereas no such association was observed among women with high resilience against negative feedback. Implications of the concept of resilience for the study of PSE are discussed.  相似文献   

16.
We examined if perfectionism and the perception of being an anxious person were associated with more negative infant temperament ratings by the mothers. 386 women (mean age = 30.08; standard deviation = 4.21) in their last trimester of pregnancy completed the Multidimensional Perfectionism Scale (MPS), the Beck Depression Inventory-II (BDI-II) and an item about their perception of being or not an anxious person. The Portuguese version of the Diagnostic Interview for Genetic Studies and the Operational Criteria Checklist for Psychotic Illness were used to generate diagnoses according to DSM-IV and ICD-10 criteria. After delivery, women completed eight items of the Difficult Infant Temperament Questionnaire (developed by our team) and filled in, again, the BDI-II and were interviewed with the DIGS. Women with depression (DSM-IV/ICD-10) and probable cases of depression using different cut-offs adjusted to Portuguese prevalence (BDI-II), in pregnancy and postpartum, were excluded. The Difficult Infant Temperament Questionnaire showed to have factorial validity and internal consistency. There was a statistically significant negative correlation between perfectionism total scale score and item 6 from the temperament scale (“is your baby irritable or fussy?”). Considering MPS 3-factor solution found for pregnancy there was also a statistically significant negative correlation between SOP and the same item. Women with low SOP differed from those with medium and high SOP in the total temperament score. Moreover, the low SOP group differed from the medium group on items three and four scores. There were no significant associations with SPP, which is the dimension more closely associated with negative outcomes. There was an association between anxiety trait status (having it or not) and scoring low, medium or high in the infant temperament scale. The proportion of anxious vs. non-anxious women presenting a high score on the infant temperament scale was higher (24.2% vs. 12.9%). Linear regressions showed that SOP (low vs. medium/high) offered a significant contribution to the prediction of total temperament scale score and items 3 and 4 scores, but a logistic regression did not confirm trait anxiety as a significant predictor of mother's infant temperament perception. Concluding, a major result concerns the fact that higher levels of adaptive perfectionism (i.e. SOP) are associated (and predict) a less negative view of their infant's temperament. These results on the effect of mother's anxiety and perfectionism on the child temperament perception might have treatment implications. As perfectionism is not always maladaptive, some of its positive features could be used to enhance women's self-efficacy/sense of parental competence in their role as mothers and positive affect towards their infants. Also, antenatal interventions aimed at minimising anxiety could help to optimise infant temperament outcomes, which could, eventually, also, lead to subsequent maternal and infant mental health better outcomes.  相似文献   

17.
ObjectiveThis study examined longitudinal relations between maternal bonding and infant temperament in the first nine months after birth.DesignOur sample consisted of 281 women, enrolled at five maternity hospitals, who completed questionnaires during the first week (T1), at six weeks (T2) and nine months postpartum (T3). Maternal bonding was assessed using the Mother-to-Infant Bonding Scale at T1 and T2 and the Postpartum Bonding Questionnaire at T3. Infant temperament was measured using the Infant Characteristics Questionnaire, completed by the mothers at T2 and T3.ResultsThe results of a path model showed a long-term effect flowing from the child to the mother, with infant temperament at T2 predicting maternal bonding at T3 over and above stability in bonding. At T3, bonding was linked more strongly to child temperament at T2 than to child temperament assessed concurrently at T3. Maternal bonding did predict infant temperament, but this was true only of bonding reported at T1 and infant temperament at T2, that is, not of bonding assessed at T2 and infant temperament at T3.ConclusionOur results indicate that maternal bonding in the first week postpartum may temporarily affect child temperament, but infant’s temperament several weeks after birth – rather than several months postpartum – plays a pervasive role in shaping the long-lasting nature of the mother-child relationship. Our findings thus seem to support the suggestion that the early postpartum weeks represent an important period in the development of maternal bonding.  相似文献   

18.
Although an important theoretical concept, little is known about the development of maternal self‐esteem. This study explores the significance of maternal cognitions, psychopathological symptoms, and child temperament in the prediction of prenatal and postnatal maternal self‐esteem. During pregnancy 162 women completed measures assessing their unhealthy core beliefs, psychopathological symptoms, and self‐esteem. At 1 year postpartum 87 of these women completed measures assessing their self‐esteem and their child's temperament. Overall maladaptive maternal core beliefs and psychopathological symptoms during pregnancy explained 19% of the variance in prenatal maternal self‐esteem. Forty‐two percent of the variance in maternal self‐esteem at 1 year could be explained by a combination of prenatal maternal self‐esteem, mental health symptoms, maternal core beliefs, and more unsociable infant temperament. Underlying maternal cognitive structures may be important in determining the development of maternal self‐esteem.  相似文献   

19.
This study examined gender differences in crying as well as associations between basic personality traits and self‐reported indices of crying. Forty‐eight men and 56 women completed the Five‐Factor Personality Inventory and the Adult Crying Inventory. Substantial gender differences were demonstrated in crying frequency and crying proneness, but not with respect to mood changes after crying. As predicted, women reported a higher frequency of crying and more proneness to cry both for negative and positive reasons. For women, all these crying indices were negatively associated with Emotional Stability. For men, only a significant negative relationship between Emotional Stability and crying for negative reasons emerged. No clear links were found between personality and mood changes after crying. Multiple regression analysis revealed a significant predictive role of gender for crying proneness, even when controlling for personality differences, but not for crying frequency. Adding personality by gender interaction terms resulted in a disappearance of the main effect of sex, while significant interactions with personality factors showed up for crying frequency and general crying proneness. It is suggested that future research on the relationship between personality and crying should focus more on the underlying mechanisms of observed relationships. Furthermore, it is recommended that future research should examine the role of different emotion regulation strategies. In addition, biological factors, temperament, upbringing measures, and socio‐demographic variables should be taken into account. Copyright © 2001 John Wiley & Sons, Ltd.  相似文献   

20.
The relationship between personality and mental health was investigated in one cohort of police trainees at a South African police academy (1145 police recruits; 648 men, 497 women). Male trainees reported less somatisation, depression, anxiety, and phobic anxiety symptoms and lower harm avoidance as well as higher persistence than female trainees. A cluster analysis based on the personality scores was used to identify three clusters with personality profiles characterized as Vulnerable, Healthy, and Intermediate profiles. Sociodemographic variables and temperament and character domain scores contributed separately and differentially to the explanation of variance in mental health symptom scores. Selection tools should be developed to identify vulnerable individuals in terms of personality characteristics during selection and prior to training, to prevent later problems with stress reactions. Additional training modules focusing on coping skills could possibly reduce vulnerability to stress in some trainees.  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