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.
BACKGROUND: Although some evidence suggests a bimodal age at onset of panic attacks, the literature comparing subjects with early versus late onset is limited. Previous work suggests that people with late-onset panic attacks may have fewer panic symptoms and exhibit less avoidance. This study sought to compare late-onset panic attacks and early-onset panic attacks with regard to (I) comorbidity, (2) health care utilization, and (3) illness behaviors and coping. METHOD: This community-based study involved interviewing randomly selected adults for the presence of DSM-III-R panic attacks. If panic attacks were confirmed, subjects were asked questions concerning panic characteristics, psychiatric comorbidity, symptom perceptions, illness attitudes, coping, and family characteristics. Subjects reporting early-onset panic (panic onset < 50 years of age) were compared with those reporting late onset (onset > or = 50 years of age). Significant univariate analyses were controlled for differences in age, panic duration, and socioeconomic status by using analysis of covariance and logistic regression. RESULTS: Subjects with late-onset panic attacks (N = 9) utilized the mental health sector less, but were more likely to present to family physicians for their worst panic. Patients with late-onset panic felt that choking and numbness more strongly disrupted function, but felt less strongly that either depersonalization or sweating disrupted function. Subjects with late-onset had fewer comorbid conditions and lower Symptom Checklist-90 scores. Late-onset subjects also had less hypochondriasis and thanatophobia while coping less through avoidance or wishful thinking. CONCLUSION: Late-onset panic attacks are associated with less mental health utilization, lower levels of comorbidity, less hypochondriasis, and a greater number of positive coping behaviors.  相似文献   

2.
Illness attitudes and coping process in subjects with panic attacks.   总被引:1,自引:0,他引:1  
Personality disorders may affect comorbidity and health care utilization in subjects with panic attacks. The purpose of this study was to identify differences in illness attitudes and behaviors, and in coping strategies in community-based subjects with panic attacks compared with controls. A community-based sample of 97 subjects meeting DSM-III-R criteria and matched controls completed the Illness Behavior Questionnaire, the Illness Attitude Scale, and the Ways of Coping Checklist. The panic group reported less healthy attitudes and behaviors. Although there were no panic-control differences in use of positive coping strategies, the panic group reported more use of negative coping strategies. The differences in illness attitudes and behaviors, and in coping strategies, may explain comorbidity and care-seeking behavior in subjects with panic attacks.  相似文献   

3.
This study examines the relative contribution of biological and psychological processes to the induction of panic attacks by a biochemical challenge agent. Panicogenic doses of caffeine were administered to 8 panic disorder (PD) patients and 11 healthy volunteers during stage 3–4 sleep, when cognitive processing is minimal and the threshold to external stimuli is high. Panic attacks were induced directly from sleep in 3 subjects and subclinical panics in an additional 3. Subjects who experienced full panic attacks spent periods of time ranging from 4 to 52 minutes in stage 2 sleep before awakening in a panic, while those who awakened in subclinical panic awakened almost directly from stage 4 sleep. PD patients experienced significantly more panic symptoms than healthy volunteers. Although limited by a small sample size, this study suggests a combined biological-psychological model of panic induction in which panic disorder patients are more biologically predisposed than healthy controls to panic symptoms but may require cognitive processing for the elaboration of a full panic attack. Depression and Anxiety 8:126–130, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

4.
Our purpose in this study was to compare the prevalence and pattern of Axis I and II comorbidities between patients with and without nocturnal panic (NP) attacks. One hundred and sixteen subjects with panic disorder (PD; according to DSM-IV criteria) were included: We assessed Axis I and II comorbidities using the Structured Clinical Interview for DSM-IV Axis I and II disorders, respectively. Of the sample, 27.6% of subjects had recurrent nocturnal panic attacks (NP group). Subjects with NP did not differ from those without in any sociodemographic or clinical characteristics. In the sample (94 subjects), 81% had at least one lifetime comorbid Axis I disorder, without significant differences between subjects with and without nocturnal panic even when considering comorbidity rates for single disorders; a trend toward significance was found for anorexia nervosa and somatization disorder, which both were more frequent among subjects with NP. Concerning Axis II disorders, 49.1% of the sample (57 subjects) met the criteria for at least one personality disorder, without significant differences between patients with and without NP. No significant differences were detected in comorbidity rates for any single Axis II personality disorder. Personality might play a relevant role in influencing treatment approaches to PD, but it does not appear to be a differential focus of concern in patients with compared to those without NP.  相似文献   

5.
BACKGROUND: Inhalation of carbon dioxide (CO(2)) has been shown to produce more anxiety in patients with panic disorder (PD) than in healthy comparison subjects or patients with most other psychiatric illnesses tested, although premenstrual dysphoric disorder (PMDD) may be an exception. Several reasons have been proposed to explain CO(2) breathing effects in PD. We examined differences in respiratory response to CO(2) breathing in 4 groups to address these issues. METHODS: Patients with PD (n = 52), healthy controls (n = 32), patients with PMDD (n = 10), and patients with major depression without panic (n = 21) were asked to breathe 5% and 7% CO(2). Continuous measures of respiratory physiological indices were made. RESULTS: Carbon dioxide breathing produced the expected increases in all 4 respiratory variables measured. More patients with PD and PMDD had panic attacks than did controls or patients with major depression. Subjects who experienced panic during 5% or 7% CO(2) inhalation had the most extreme increases regardless of diagnostic group. Among patients with PD, baseline end-tidal carbon dioxide levels were significantly lower in those who subsequently had a panic attack during 5% CO(2) breathing than those who did not. CONCLUSIONS: Although CO(2) breathing causes a higher rate of panic attacks in patients with PD than other groups (except PMDD), the physiological features of a panic attack appear similar across groups. Once a panic attack is triggered, minute ventilation and respiratory rate increase regardless of whether the subject carries a PD diagnosis. These findings are compatible with preclinical fear conditioning models of anxiogenesis.  相似文献   

6.
Although recent diagnostic systems support the distinctiveness of panic disorder (PD) and somatization disorder, a high level of comorbidity of these two diagnoses has been reported, indicating a need for investigations with external validators. One hundred fifty-nine outpatients with DSM-III-R PD and 76 surgical controls were screened for lifetime presence of DSM-III-R somatization disorder, and the risks for some types of psychiatric disorders in their families were computed. In our sample, 23% of women and 5% of men with PD also had DSM-III-R somatization disorder. Women patients with PD plus somatization disorder did not differ from women with PD only in age at onset of panic, agoraphobia, childhood history of separation anxiety, or lifetime diagnoses of other disorders. Familial risks for PD, PD-agoraphobia, and alcohol dependence were significantly higher for families of women with PD and women with PD plus somatization disorder than for controls. The familial risks for antisocial personality (ASP) disorder (a familial indicator for the somatization disorder spectrum of liability, phenomenologically independent from both PD and somatization disorder) were significantly higher for families of women with PD plus somatization disorder than for families of women with PD only or for controls. Application of DSM-IV criteria for somatization disorder substantially decreased the comorbidity with PD. Our data suggest that somatization disorder is not simply a form of PD, and that the two disorders may coexist in the same subject without sharing a common genetic diathesis. Compared with DSM-III-R, DSM-IV criteria for somatization disorder appear to be simpler in structure and of less complicated application.  相似文献   

7.
In order to examine the validity of the distinction between generalized anxiety disorder (GAD) and panic disorder (PD) we compared 41 subjects with GAD and 71 subjects with PD. The GAD subjects had never had panic attacks. In contrast to the symptom profile in PD subjects suggestive of autonomic hyperactivity, GAD subjects had a symptom pattern indicative of central nervous system hyperarousal. Also, subjects with GAD had an earlier, more gradual onset of illness. In terms of coexisting syndromes, GAD subjects more often had simple phobias, whereas PD subjects more commonly reported depersonalization and agoraphobia. GAD subjects more frequently had first-degree relatives with GAD, whereas PD subjects more frequently had relatives with PD. A variety of measures indicated that our GAD subjects had a milder illness than those with PD. Also, fewer GAD subjects gave histories of major depression than did PD subjects. Among GAD subjects, coexisting major depression was associated with simple phobia and thyroid disorders and among PD subjects, comorbid depression was associated with social phobia and hypertension. Our findings indicate that the separation of GAD from PD is a valid one. They also indicate that, within disorders, unique patterns of comorbidity may exist that are important both clinically and theoretically.  相似文献   

8.
BACKGROUND: Panic disorder and agoraphobia are closely linked. There are indications that uncontrolled panic attacks often lead to the rapid development of phobic avoidance, but our ability to predict which individuals with panic will develop avoidance has been limited. The purpose of this study was to identify independent predictors of the development of phobic avoidance and the time course of that development. METHOD: We conducted a secondary analysis of survey data from the community-based Panic Attack Care-Seeking Threshold Study. The presence of panic attacks was confirmed in 97 randomly selected adults from randomly selected households screened using the Structured Clinical Interview of DSM-III-R (SCID). The presence of limited and extensive phobic avoidance was measured using the SCID, while rapidity of development (lag time) was measured as the difference between onset of panic and onset of avoidance. Predictors considered included panic characteristics, psychiatric comorbidity, cognitive appraisal, family characteristics, illness attitudes, symptom perceptions, and coping style. RESULTS: Thirty-six subjects (37%) had at least mild phobic avoidance, with 81% (N = 29) of those developing the avoidance less than 1 year after the onset of panic attacks. The development of phobic avoidance was associated with the presence of panic disorder (beta = 1.36), the number of comorbid psychiatric disorders (beta = 0.69), and the number of family members and/or friends available to discuss health concerns (beta = 0.87). Further progression to agoraphobia was predicted by the presence of depersonalization during panic attacks (beta = 0.50). Rapid onset of avoidance (panic avoidance lag time < 1 year) was predicted by the perception that depersonalization is a life-threatening symptom (beta = 1.56). CONCLUSION: The development of phobic avoidance is closely linked to panic attacks and often develops soon after panic onset. Full-blown panic disorder and psychiatric comorbidity are important in this development. Depersonalization is also key to the development of avoidance and the rapidity of the development.  相似文献   

9.
The diagnostic utility of lactate sensitivity in panic disorder   总被引:2,自引:0,他引:2  
Lactate infusion is the most extensively studied of the pharmacological challenge tests in panic disorder. We assessed the value of this test in the diagnosis and subtyping of panic in clinical and research settings. Analysis of lactate infusion studies to date suggests that patients with panic attacks are significantly more sensitive to lactate than are healthy controls or patients with other psychiatric disorders without panic attacks. However, the usefulness of lactate infusion is limited by the lack of standardized, objective criteria for lactate-induced panic and uncertainty as to the sensitivity and specificity of the test for current, clinically significant panic attacks. Except in rare cases, the clinical history is likely to be of more value than lactate response in diagnosing panic disorder. Determination of the role of the test in subtyping patients with panic disorder awaits further study of the diagnostic, prognostic, genetic, and pathophysiologic significance of lactate sensitivity.  相似文献   

10.
Immunity, major depression, and panic disorder comorbidity.   总被引:4,自引:0,他引:4  
Because recent research reports indicated clinical and biological differences in major depression with and without comorbid Diagnostic and Statistical Manual of Mental Disorders (DSM-III-R) panic disorder, and as altered immune measures were reported in selected subgroups of depressive patients, we investigated 51 pairs of major depressive episode (MDE) subjects, and gender- and age-matched healthy controls in order to determine if T lymphocytes number and function abnormalities were associated with Panic Disorder comorbidty. We found that those MDE subjects with DSM-III-R panic disorder (PD) had greater numbers of T cells (p less than 0.05) and PHA mitogen (p less than 0.05) responses than depressive patients without PD, as well as increased phytohemagglutinin (PHA) (p less than 0.05) concanavalin A (ConA) (p less than 0.02) mitogen responses compared to their controls. These data suggest that panic disorder comorbidity significantly contributes to the variance of immunologic parameters in major depression and has to be carefully assessed within psychoimmunological studies of psychiatric patients with affective disorders.  相似文献   

11.
The present study examined the impact of comorbid major depressive disorder (MDD) on psychiatric morbidity, panic symptomatology and frequency of other comorbid psychiatric conditions in subjects with panic disorder (PD). Four hundred thirty-seven patients with PD were evaluated at intake as part of a multicenter longitudinal study of anxiety disorders; 113 of these patients were also in an episode of MDD. Patients were diagnosed by DSM-III-R criteria utilizing structured clinical interviews. The 113 PD/MDD patients were compared with the 324 remaining PD subjects regarding panic symptoms at intake, sociodemographic, quality of life and psychiatric morbidity variables. Differences in frequency of other comorbid Axis I psychiatric disorders were assessed at intake; personality disorders were evaluated twelve months after intake. The results revealed that PD/MDD patients exhibit increased morbidity and decreased psychosocial functioning as compared to PD patients. Personality disorders were more prevalent in the PD/MDD group at six month follow-up assessment; the PD/MDD group also had an increased frequency of posttraumatic stress disorder (PTSD) and more comorbid Axis I anxiety disorders as compared to the PD group. The total number and frequency of panic symptoms was highly consistent between the two patient groups. Depression and Anxiety 5:12–20, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

12.
Although several studies have indicated that a substantial portion of alcoholics have an anxiety disorder, relatively little information exists specifically regarding panic disorders. In addition, prior studies have been marred by the absence of appropriate contrast groups. The present investigation compared the lifetime prevalence of panic attacks and panic-related disorders diagnosed according to DSM-III criteria in a group of 79 alcohol-dependent patients, 64 depressed patients, and 70 nonclinical subjects. Panic attacks, panic disorder, and agoraphobia with panic attacks were more prevalent in the alcohol-dependent and depressed samples than among nonclinical subjects. Men in both clinical samples were more likely than women to have had nonagoraphobic panic disorder, but male alcoholics were less likely to have developed agoraphobia than were female alcoholics or depressed patients of either sex. No consistent chronological relationship between onset of panic attacks and alcohol abuse was found. Results indicate that there is an unusually high prevalence of panic attacks and panic-related disorders among alcoholics, but comparable prevalence rates can be found in depressed and perhaps some other psychiatric populations. Implications for the assessment and treatment of alcohol-dependent and panic-disordered patients are discussed.  相似文献   

13.
The frequency and severity of separation anxiety for subjects with panic disorder and major depression was compared with that for normal controls. The subjects were diagnosed according to DSM-III criteria. Each subject completed a questionnaire consisting of 9 items derived from DSM-III criteria for separation anxiety disorder. The incidence of separation anxiety and its severity were significantly higher for the panic disorder subjects than for normal controls but there was no significant difference between depressed and panic disorder subjects. Panic disorder subjects with a history of separation anxiety disorder had a significantly earlier onset of panic attacks.  相似文献   

14.
Eighty-nine subjects with panic disorder, who had been naturalistically treated, and 46 nonanxious controls were followed up after 3 years. Although they remained symptomatic, most subjects with panic disorder reported relatively little distress or social maladjustment. The course of panic disorder was characterized by fluctuating anxiety and depressive symptoms. Panic subtypes (uncomplicated, limited phobic avoidance, and extensive phobic avoidance) and Axis I and II comorbidity (major depression and personality disorders) were highly predictive of symptoms and social adjustment after 3 years. Abnormal personality was, in fact, the strongest predictor of social maladjustment in both subjects with panic disorder and controls. The results showed that while panic disorder has a favorable outcome, the illness is a chronic one that may require continuing treatment. They also show that subtypes and comorbid disturbances are important predictors of outcome.  相似文献   

15.
OBJECTIVE: The authors compared young and older adults with panic disorder (PD) to investigate differences in panic-associated phenomenology, psychiatric comorbidity, and risk factors. METHOD: Patients in the older group (age 60 and above) were further subdivided into early- and late-onset groups and compared. Phenomenology (number of panic symptoms, severity of anxiety, physiological symptoms, panic-associated cognitions, and overall severity of PD); comorbidity (depressive and anxiety disorders); and risk factors (family history of anxiety and life stressors) were assessed in 167 outpatients with PD. RESULTS: Older patients reported fewer panic symptoms, less anxiety and arousal, less severe PD, lower levels of depression, and higher levels of functioning. Furthermore, within the older-patient group, late-onset patients were found to report less distress during panic attacks in relation to body sensations and panic-related cognitions and emotions. Multiple-regression analysis of the entire sample showed that chronological age and age at onset of PD distinctly predicted different domains of panic phenomenology. CONCLUSION: PD was consistently less severe in older patients across multiple domains, and a later age at onset was associated with less distress due to body sensations, cognitions, and emotions during panic attacks.  相似文献   

16.
By using data from the Bremer Adolescent Study, this report presents findings on the frequency, comorbidity, and psychosocial impairment of panic disorder and panic attacks among 1,035 adolescents. The adolescents were randomly selected from 36 schools in the province of Bremen, Germany. Panic disorder and other psychiatric disorders were coded based on DSM-IV criteria using the computerized-assisted personal interview of the Munich version of the Composite International Diagnostic Interview. Panic disorder occurred rather rare, with only 0.5% of all the adolescents met the DSM-IV criteria for this disorder sometimes in their live. Panic attack occurred more frequently, with 18% of the adolescents reported having had at least one panic attack. Slightly more girls than boys had panic attack and panic disorder. The occurrence of panic attack and panic disorder were the greatest among the 14–15 year olds. The experience of having a panic attack was associated with a number of problems, the most frequent being avoiding the situation for fear of having another attack. Four most common symptoms associated with a panic attack were that of palpitations, trembling/shaking, nausea or abdominal distress, and chills or hot flushes. Panic disorder comorbid highly with other psychiatric disorders covered in our study, especially with that of major depression. Among those with a panic disorder, about 40% of them were severely impaired during the worst episode of their illness. Only one out of five adolescents with panic disorder sought professional help for emotional and psychiatric problems. The implication of our findings for research and clinical practice are discussed. Depression and Anxiety 9:19–26, 1999. © 1999 Wiley-Liss, Inc.  相似文献   

17.
Intravenous lactate infusion and 60 mg oral fenfluramine challenges were administered to 26 patients with panic disorder (PD) and 12 age- and sex-matched control subjects. PD patients had significantly greater anxiety responses to either challenge than controls. When a panic attack frequency index scale of 1-4 was used, PD patients with more recent and more frequent spontaneous panic attacks scored higher on either challenge. There was a significant correlation between increasing panic attack frequency and greater anxiogenic responses to lactate and fenfluramine. Nine of 12 patients (75%) with high attack frequency (defined as having one or more panic attacks per week) reacted positively to both challenges in contrast to 0 of 14 PD patients with low frequency (less than or equal to 1 attack/month). The findings suggest that the current heightened anticipatory state of the patient (influenced by recent spontaneous panic attacks) rather than putative underlying trait factors predominates in the evocation of experimentally induced anxiety reactions. Future studies must consider the frequency of recent spontaneous panic attacks in the evaluation of anxiogenic reactivity to provocative stimuli.  相似文献   

18.
19.
Urinary pH was evaluated in panic disorder (PD) patients compared with both psychiatric and healthy control subjects. Fourteen PD patients, eight major depressive disorder (MDD) patients, and 14 healthy control (HC) subjects were examined. All patients were drug-free and met DSM-IV diagnostic criteria. The PD patients had lower urinary pH and higher levels of anxiety than both MDD and HC subjects. Additionally, urinary pH inversely correlated with anxiety levels. Although preliminary, these findings suggest that PD patients have lower urinary pH than MDD and HC subjects. Future studies that simultaneously examine both urinary and blood pH in larger numbers of PD patients and patients with other anxiety disorders, before and after treatment, need to be conducted.  相似文献   

20.
The evidence supporting the existence of panic disorder as a distinct clinical entity is critically examined, as are the current criteria for panic disorder in DSM-III. It is argued that the current definition of a panic attack is imprecise and that the borders and overlap of panic disorder with other psychiatric disorders raise broader questions as to what is meant by a distinct psychiatric disorder. DSM-III "panic disorder" defines an ideal type that may be more relevant for research purposes than clinical. In defining fairly homogeneous "pure" cases, it overlooks the prevalence and importance of atypical "mixed" and subsyndromal cases.  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