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.
We investigated Buss and Plomin’s Emotionality Activity Sociability (EAS) Temperament Survey for Children, used to assess temperament. Temperament is believed to comprise stable traits that change little over time. We examined stability of EAS temperament in the Avon Longitudinal Study of Parents and Children, in which 7429 mothers completed the EAS when their children were 3, 5 and 6 years old. Factor analysis was conducted at each time point, and stability over time was assessed using correlation and mixed effects regression modelling, accounting for differences within and between individuals. A four-factor model provided the best fit, with 19–20 of the 20 items loading onto the dimensions predicted by Buss and Plomin. Test–retest correlations ranged from 0.52–0.64 (3–5 years), 0.59–0.74 (5–6 years) and 0.46–0.58 (3–6 years). Mixed effects regression modelling suggested good stability over time: differences between, rather than within, individuals accounted for between 69% and 82% of the variance. This study demonstrates the stability of temperament over time, a vital pre-requisite to investigating childhood temperament as a predictor of outcomes.  相似文献   

2.
The present work analyzes the relationships between the dimensions of temperament and the exteriorized emotions of aggression and anger. Temperament was assessed by mothers using the Dimensions of Temperament Survey‐Revised, while aggression and anger were self‐reported by the children using the Scale of physical and Verbal Aggression and the State–Trait Anger Expression Inventory for Children. The sample studied was made up of 293 children (49.83% boys; 50.17% girls) with a mean age of 11.13 years. The results showed that temperamental difficulties give rise to exteriorized emotions, especially anger. Predictive values of temperament on aggression and anger ranged from 1% to 7% of explained variance. Aggr. Behav. 32:207–215, 2006. © 2006 Wiley‐Liss, Inc.  相似文献   

3.
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   

4.
Temperament and Happiness in Children   总被引:1,自引:0,他引:1  
The relation between temperament and happiness was assessed in 311 children aged 9–12. Parents rated their children’s temperament using the Emotionality, Activity, and Sociability Temperament Survey (EAS) and rated their children’s happiness. Children rated their own temperament using the EAS and the Piers–Harris Self Concept Scale for Children Second Edition, and they rated their own happiness using a single-item measure, the Oxford Happiness Questionnaire Short Form, and the Subjective Happiness Scale. Parents’ and children’s temperament ratings conformed to the four factor structure proposed by Buss and Plomin(Temperament: Early developing personality traits, Lawrence Erlbaum Associates, Hillsdale, 1984) supporting the use of children’s self-reports as an additional measure of temperament. Temperament accounted for between 9 and 29% of the variance in children’s happiness depending on the measures. Children who were more social and active, and less shy, emotional, and anxious were happier. These results parallel the well-established relation between happiness and personality in adults; temperament traits akin to extraversion (Sociability) were positively associated with happiness whereas traits akin to neuroticism (Emotionality) were negatively associated with happiness. Additionally, children who were rated higher in the temperament trait Activity were happier.  相似文献   

5.
The relation between temperament and happiness was assessed in a sample of 441 children aged 7–14 years drawn from a population in Northern India. Parents assessed their children’s happiness and rated their children’s temperament using the Emotionality, Activity, and Sociability Temperament Survey (EAS). Children self-reported their own happiness using a single-item measure, the Oxford Happiness Scale Short Form, and the Subjective Happiness Scale. Parents’ temperament ratings conformed to the four factor structure proposed by Buss and Plomin (Temperament: early developing personality traits. Lawrence Erlbaum Associates, Hillsdale, 1984): Emotionality, Activity, Sociability, and Shyness. Temperament accounted for between 4 and 11% of the variance in children’s happiness depending on the measures. Children who were more social and active, and less shy, were happier. This result parallels the well-established relation between happiness and personality in adults and is similar to recent research on happiness and temperament in children; temperament traits akin to extraversion were positively associated with happiness. However, despite that neuroticism and its temperament counterpart are strongly and consistently linked to happiness in adults, the relation between happiness and the temperament trait associated with neuroticism (i.e., Emotionality) was weak. This suggests that the relations between temperament and happiness in children may not completely generalize across cultures.  相似文献   

6.
Individual differences among adults have generally been conceptualized in terms of personality theory and traits. In contrast, individual differences among very young children (birth to kindergarten) have generally been conceptualized in terms of temperament theory and traits. The present study compares and contrasts measures of temperament and personality in a sample of preschool children. Temperament traits were assessed with a well‐established measure (the Rothbart CBQ), and a new preschool rating instrument was used to assess personality traits from the five‐factor framework (M5‐PS). Indeed, a key purpose of this study was to further the development of the M5‐PS. Data were gathered on 122 preschool children who were rated by their teachers. Significant correlations were found between the temperament trait Surgency and the personality trait Extraversion, between the temperament trait Negative Affect and the personality trait Neuroticism, and between the temperament trait Effortful Control and the personality trait Conscientiousness. The overall pattern of correlational data suggests that individual differences in preschool children can be adequately described using the five‐factor theory, and that this framework may effectively subsume traditional theories of temperament. Preliminary support for the reliability and validity of the M5‐PS is offered. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

7.
Background. Student's temperament plays a significant role in teacher's perception of the student's learning style, educational competence (EC), and teachability. Hence, temperament contributes to student's academic achievement and teacher's subjective ratings of school grades. However, little is known about the effect of gender and teacher's age on this association. Aims. We examined the effect of teacher's and student's gender and teacher's age on teacher‐perceived temperament, EC, and teachability, and whether there is significant same gender or different gender association between teachers and students in this relationship. Sample. The participants were population‐based sample of 3,212 Finnish adolescents (M= 15.1 years) and 221 subject teachers. Methods. Temperament was assessed with Temperament Assessment Battery for Children – Revised and Revised Dimensions of Temperament Survey batteries and EC with three subscales covering Cognitive ability, Motivation, and Maturity. Data were analyzed with multi‐level modelling. Results. Teachers perceived boys’ temperament and EC more negatively than girls’. However, the differences between boys and girls were not as large when perceived by male teachers, as they were when perceived by female teachers. Males perceived boys more positively and more capable in EC and teachability than females. They were also stricter regarding their perceptions of girls’ traits. With increasing age, males perceived boys’ inhibition as higher and mood lower. Generally, the older the teacher, the more mature he/she perceived the student. Conclusions. Teachers’ ratings varied systematically by their gender and age, and by students’ gender. This bias may have an effect on school grades and needs be taken into consideration in teacher education.  相似文献   

8.
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   

9.
Conventionally, individual differences have been assessed using temperament measures for infants and children, and personality measures for adults. We chose to explore both temperament and personality to see whether a convergence exists specifically during adolescence. A sample of 225 adolescents completed Rothbart's Early Adolescent Temperament Questionnaire–Revised (EATQ–R), a 4-factor temperament scale, and the HEXACO Personality Inventory–Revised (HEXACO PI–R), a 6-factor personality scale. As hypothesized, we found significant relations between the 2 measures. However, there were some important differences between the 2 measures regarding Honesty–Humility, Openness, and Frustration that highlight the unique contributions of both instruments to understanding and measuring adolescent individual differences. As there is a relatively scant history of measuring temperament or personality in adolescence, it is sometimes difficult for researchers to decide which instrument is most appropriate. The results reported here suggest that either the EATQ–R or the HEXACO PI–R could be appropriate, depending on the specific research questions being asked.  相似文献   

10.
This study had three objectives: (1) to assess the relationship between the single nucleotide polymorphism (SNP) rs1789891 in the alcohol dehydrogenase gene cluster and alcohol dependence and affective disorders; (2) to assess the differences in the Regulative Theory of Temperament (RTT) traits between an alcohol dependent group, an affective disorders group, and a healthy group; and (3) to assess the relationship between rs1789891 and temperament traits in a healthy group, taking into account the interaction of genotype and sex. The SNP rs1789891 was genotyped in a group of 194 alcohol dependent men, aged 21 to 71 years; 137 patients with affective disorders, including 51 males and 86 females, aged 19 to 85 years; and a group of 207 healthy individuals, including 89 males and 118 females, aged 18 to 71 years. Temperament traits (briskness, perseveration, sensory sensitivity, emotional reactivity, endurance, and activity) were assessed in all groups using the Formal Characteristics of Behaviour‐Temperament Inventory. The comparative analysis of genotypic frequencies showed no significant differences between patients with alcoholism or affective disorders and those in the control group. Alcohol dependent men and the affective disorder group were characterised by higher levels of emotional reactivity (p‐value 1.4e‐5 and 9.84e‐7, respectively) and lower levels of briskness, sensory sensitivity, endurance, and activity (p‐value from 3.76e‐8 to 0.012) when compared to the healthy group. The rs1789891 polymorphism was associated with briskness (= 0.02), sensory sensitivity (= 0.036), and activity (= 0.049). None of the results were statistically significant after Bonferroni correction.  相似文献   

11.
This study examined the relationships between the Five-Factor-Model (FFM) personality dimensions (Extraversion, Agreeableness, Conscientiousness, Emotional Stability, Openness/Autonomy), the Pavlovian temperament variables (Strength of Excitation, Strength of Inhibition, Mobility), and fatigue. We expected that all these person characteristics would be negatively associated with fatigue. In a survey among persons working at least 20 h a week (N=765), respondents completed a personality questionnaire (the Five-Factor Personality Inventory), a temperament scale (the Pavlov Temperament Survey), and two fatigue questionnaires (the Checklist Individual Strength-20 and the Emotional Exhaustion scale of the Maslach Burnout Inventory). Results indicated that high scores on Autonomy, and low scores on Extraversion, Conscientiousness, Emotional Stability, and Strength of Excitation were, in various combinations, predictors of higher fatigue scores. Agreeableness, Strength of Inhibition, and Mobility did not play a role in the prediction of any fatigue score. In conclusion, personality and temperament dimensions explained significant proportions of the variance in fatigue. It is stated that there is an urgent need for longitudinal studies in order to examine the predictive value of the personality dimensions over time.  相似文献   

12.
Positive parenting has been related both to lower cortisol reactivity and more adaptive temperament traits in children, whereas elevated cortisol reactivity may be related to maladaptive temperament traits, such as higher negative emotionality (NE) and lower positive emotionality (PE). However, no studies have examined whether hypothalamic‐pituitary‐adrenal axis activity, as measured by cortisol reactivity, moderates the effect of the quality of the parent–child relationship on changes in temperament in early childhood. In this study, 126 3‐year‐olds were administered the Laboratory Temperament Assessment Battery (Lab‐TAB; Goldsmith et al., 1995) as a measure of temperamental NE and PE. Salivary cortisol was collected from the child at 4 time points during this task. The primary parent and the child completed the Teaching Tasks battery (Egeland et al., 1995), from which the quality of the relationship was coded. At age 6, children completed the Lab‐TAB again. From age 3 to 6, adjusting for age 3 PE or NE, a better quality relationship with their primary parent predicted decreases in NE for children with elevated cortisol reactivity and predicted increases in PE for children with low cortisol reactivity. Results have implications for our understanding of the interaction of biological stress systems and the parent–child relationship in the development of temperament in childhood.  相似文献   

13.
Studies of gender differences using primarily young individuals show that males, on average, perform better than females in physical activities but worse than females on tests of verbal abilities. There is however a controversy about the existence of these sex differences in adulthood. Our study used 1271 participants from four cultural backgrounds (Chinese, multi‐generation Canadians, Indu‐Canadians, and European‐Canadians) divided in five age groups. We measured sex differences in the time required for participants to complete a lexical task experiment, and also assessed their verbal tempo and physical endurance using a validated temperament test (Structure of Temperament Questionnaire). We found a significant female advantage in time on the lexical task and on the temperament scale of social–verbal tempo, and a male advantage on the temperament scale of physical endurance. These sex differences, however, were more pronounced in young age groups (17–24), fading in older groups. This “middle age–middle sex” phenomenon suggests that sex differences in these two types of abilities observed in younger groups might be “a matter of age,” and should not be attributed to gender in general. A one‐dimensional approach to sex differences (common in meta‐analytic studies) therefore overlooks a possible interaction of sex differences with age.  相似文献   

14.
A genetically informed longitudinal cross‐lagged model was applied to twin data to explore etiological links between difficult temperament and negative parenting in early childhood. The sample comprised 313 monozygotic (MZ) and dizygotic (DZ) twin pairs. Difficult temperament and negative parenting were assessed at ages 2 and 3 using parent ratings. Both constructs were interrelated within and across age (rs .34–.47) and showed substantial stability (rs .65–.68). Difficult temperament and negative parenting were influenced by genetic and environmental factors at ages 2 and 3. The genetic and nonshared environmental correlations (rs .21–.76) at both ages suggest overlap at the level of etiology between the phenotypes. Significant bidirectional associations between difficult temperament and negative parenting were found. The cross‐lagged association from difficult temperament at age 2 to negative parenting at age 3 and from negative parenting at age 2 and difficult temperament at age 3 were due to genetic, shared environmental, and nonshared environmental factors. Substantial novel genetic and nonshared environmental influences emerged at age 3 and suggest change in the etiology of these constructs over time.  相似文献   

15.
Although the terms temperament and personality are often used interchangeably in the literature and clearly are conceptually related, there is little empirical data to illuminate their relationship. In this exploratory study we measured temperament (using the Dimensions of Temperament Survey), and personality (using the Eysenck Personality Questionnaire), in young adolescents and in adults. Some age and sex differences in relationships between the measures and their correlations were found. The temperament dimension of adaptability was clearly related to Eysenck's Extraversion factor and reactivity to Neuroticism in females. More substantial relationships emerged in the adult sample than in the adolescent group, Our data, combined with that of a 1984 Australian study using different techniques, provides substantial support for measurable overlap between some dimensions of temperament derived largely from the developmental literature, and Eysenck's personality theory.  相似文献   

16.
In this study, we analyzed the relationship between temperament and personality factors and depression in children and adolescents. Temperament was assessed with the Dimensions of Temperament Survey Revised (DOTS-R), personality with the Big Five Questionnaire-Children (BFQ-C), and depressive symptomatology with the Children's Depression Inventory (CDI). The sample was made up of 535 participants (274 boys and 261 girls), aged 8 to 15 years. Results show that temperament and personality are significantly related to depressive symptomatology in children and adolescents. Those with difficult temperaments showed more depressive symptomatology, as did those with high levels of emotional instability or low levels of extraversion, openness, agreeableness or conscientiousness. Multiple regression analyses revealed greater relevance of personality variables than of temperament variables.  相似文献   

17.
This article analyzes the differences between an activity-specific temperament model and the Big Five personality model using the Structure of Temperament Questionnaire--Compact (STQ-77). The STQ-77 has 3 emotionality scales and 9 scales assessing 3 dynamic aspects (arousal, lability, and sensory sensitivity) in 3 areas of activity (physical, verbal-social, and mental). The results of administration of the Russian STQ-77, NEO-FFI, and SSS-V to 174 Russian participants showed how components of temperament can represent the traits described in the Big Five model. The confirmatory factor analysis of the English STQ-77 and the results of a study involving a prolonged word classification task with 221 Canadian participants showed the benefits of the activity-specific approach, separating temperament traits in three areas of activity. Such specificity of temperament traits differentiates them from personality traits.  相似文献   

18.
气质的遗传因素:基因多态性研究   总被引:1,自引:0,他引:1       下载免费PDF全文
气质是婴儿出生以后最早表现出来的一种人格特征,而人格特征的形成具有一定的生物学基础。气质作为人格特征的组成单元,与人格特征享有共同的生物过程。因此,与成年人人格特征相关的基因也有可能影响儿童气质。本文在成年人的人格特征研究的基础上,针对早期儿童发展的特点,讨论了可能影响早期儿童气质的候选基因。研究成人的人格特征与基因的相关性不能排除环境因素的影响,本文强调了对早期儿童气质与基因多态性之间的关系进行直接研究的重要性。  相似文献   

19.
Temperament refers to individual differences in reactivity and self-regulation and is influenced by genetic and experiential variation and maturation. Temperament reflects biologically based individual differences that emerge in early life and remain relatively stable thereafter. Given the growing interest in cultural variation in infant temperament, this study examined the temperament of 12-month-old children in Chile and the US. The aims were to validate a version of the Infant Behavior Questionnaire – Revised – Very Short Form in Spanish for Chile and to compare Chilean and US infants’ temperament. For the first aim, 150 Chilean infants aged 10–15 months were assessed, and 73 US infants aged 10–15 months were examined for the second aim. The children’s parents completed a demographic questionnaire and the IBQ-R-VSF, which measures three dimensions of temperament: Surgency, Negative Affectivity, and Effortful Control. The reliability of each dimension for the Chilean sample was between 0.70 and 0.75, and significant differences between Chilean and US infants emerged. Parents of Chilean infants reported higher levels of Effortful Control, whereas US parents reported that their infants exhibited higher levels of Negative Affectivity. A relationship between parents’ higher educational level and infants’ higher levels of Surgency was found for both countries. No gender or age differences were observed for any of the three temperament dimensions. These results and their implications for cultural studies are discussed.  相似文献   

20.
ObjectivesThe aim of the present study was to investigate the associations among temperament traits postulated by the Regulative Theory of Temperament (RTT), posttraumatic stress disorder symptoms (PTSD symptoms), emotion regulation strategies, and affect in the 280 motor vehicle survivors (MVA).MethodsTemperament was measured with the Formal Characteristics of Behaviour–Temperament Inventory (FCB-TI), the level of posttraumatic stress disorder symptoms was assessed by the PTSD Clinical Inventory (PTSD-C), emotion regulation was tested with the Polish adaptation of the Inventory of Cognitive Affect Regulation Strategies (ICARUS), and affect was evaluated by the Polish version of the Positive and Negative Affect Scale (PANAS).ResultsGreater emotional reactivity was associated with grater negative affect (also by maladaptive regulation) and lower positive affect whereas greater activity was linked to grater positive affect (also via adaptive regulation). Furthermore, greater PTSD symptoms were related to greater negative affect (also through maladaptive regulation) and lower positive affect. However, PTSD symptoms were not linked to adaptive regulation strategies.ConclusionThe findings significantly extends our current knowledge on the associations among temperament traits, PTSD symptoms, emotion regulation strategies, and affect in the motor vehicle survivors.  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