Journal articles on the topic 'Fatigue research'

To see the other types of publications on this topic, follow the link: Fatigue research.

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Fatigue research.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Manzullo, Ellen F., and Carmen P. Escalante. "Research into fatigue." Hematology/Oncology Clinics of North America 16, no. 3 (June 2002): 619–28. http://dx.doi.org/10.1016/s0889-8588(02)00012-6.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Qiu, Xingwen, Haishan Yin, and Qicheng Xing. "Research Progress on Fatigue Life of Rubber Materials." Polymers 14, no. 21 (October 28, 2022): 4592. http://dx.doi.org/10.3390/polym14214592.

Full text
Abstract:
Rubber products will be fatigued when subjected to alternating loads, and working in harsh environments will worsen the fatigue performance, which will directly affect the service life of such products. Environmental factors have a great influence on rubber materials, including temperature, humidity, ozone, etc., all of which will affect rubber’s properties and among which temperature is the most important. Different rubber materials have different sensitivity to the environment, and at the same time, their own structures are different, and their bonding degree with fillers is also different, so their fatigue lives are also different. Therefore, there are generally two methods to study the fatigue life of rubber materials, namely the crack initiation method and the crack propagation method. In this paper, the research status of rubber fatigue is summarized from three aspects: research methods of rubber fatigue, factors affecting fatigue life and crack section. The effects of mechanical conditions, rubber composition and environmental factors on rubber fatigue are expounded in detail. The section of rubber fatigue cracking is expounded from macroscopic and microscopic perspectives, and a future development direction is given in order to provide reference for the research and analysis of rubber fatigue and rubber service life maximization.
APA, Harvard, Vancouver, ISO, and other styles
3

Chou, Yun-Jen, Kord M. Kober, Ching-Hua Kuo, Kun-Huei Yeh, Tien-Chueh Kuo, Yufeng J. Tseng, Christine Miaskowski, Jin-Tung Liang, and Shiow-Ching Shun. "A Pilot Study of Metabolomic Pathways Associated With Fatigue in Survivors of Colorectal Cancer." Biological Research For Nursing 23, no. 1 (July 22, 2020): 42–49. http://dx.doi.org/10.1177/1099800420942586.

Full text
Abstract:
Background: Over 30% of cancer survivors experience chronic fatigue. An alteration in energy metabolism is one of the hypothesized mechanisms for cancer-related fatigue (CRF). No studies have evaluated for changes in metabolic profiles in cancer survivors with CRF. The purpose of this pilot study was to evaluate for differences in metabolic profiles between fatigued and non-fatigued survivors of colorectal cancer (CRC). Methods: Survivors were recruited from the surgical outpatient department and the oncology clinic of a medical center in northern Taiwan. Fatigue was assessed using the Fatigue Symptom Inventory. Fasting blood samples were collected on the day the fatigue questionnaire was completed. Metabolomic profile analysis was performed using non-targeted, liquid chromatography/time-of-flight mass spectrometry. Fold change analyses, t-tests, and pathway analyses were performed to identify differences in metabolomic profiles between the fatigued and non-fatigued survivors. Results: Of the 56 CRC survivors in this study, 28.6% (n = 16) were in the fatigue group. Statistically significant differences in carnitine, L-norleucine, pyroglutamic acid, pyrrolidonecarboxylic acid, spermine, hydroxyoctanoic acid, and paraxanthine were found between the two fatigue groups. In addition, two pathways were enriched for these metabolites (i.e., glutathione metabolism, D-glutamine and D-glutamate metabolism). Conclusions: Findings from this pilot study provide preliminary evidence that two pathways that are involved with the regulation of ATP production and cellular energy (i.e., glutathione metabolism, D-glutamine and D-glutamate metabolism) are associated with fatigue in CRC survivors. If these findings are confirmed, they may provide new therapeutic targets to decrease fatigue in cancer survivors.
APA, Harvard, Vancouver, ISO, and other styles
4

Wang, Xian Jun, and Xiang Hui Yuan. "Research of Infrared Video Fatigue Detection Based on SoPC." Applied Mechanics and Materials 411-414 (September 2013): 1488–94. http://dx.doi.org/10.4028/www.scientific.net/amm.411-414.1488.

Full text
Abstract:
A new high-speed infrared video-based fatigue detection system was developed using a system on a programmable chip (SoPC) in this study. Based on the limitations of PERCLOS, we merged the eyes and mouth fatigue related characteristics to improve detection accuracy, used a Difference of Gaussian (DoG) filter and Ada-boosting algorithm to implement driver fatigue detection based on multi-feature fusion. The detection system was produced using FPGA with a parallel processing structure and pipeline technology. This system is innovative and it can detect fatigued states efficiently and rapidly.
APA, Harvard, Vancouver, ISO, and other styles
5

Prinsen, Hetty, Jolanda de Vries, Foekje Stelma, Sasja Mulder, Carla Van Herpen, Jan Willem Leer, Lammy Elving, Gijs Bleijenberg, and Hanneke W. M. Van Laarhoven. "Humoral and cellular immune response after influenza vaccination in patients with postcancer fatigue and patients with chronic fatigue syndrome." Journal of Clinical Oncology 30, no. 15_suppl (May 20, 2012): 9070. http://dx.doi.org/10.1200/jco.2012.30.15_suppl.9070.

Full text
Abstract:
9070 Background: Postcancer fatigue (PCF) is a frequently occurring problem, impairing quality of life. Patients with chronic fatigue syndrome (CFS) also suffer from severe fatigue symptoms. We hypothesized that in fatigued patients (PCF and CFS) alterations in immune response could explain fatigue symptoms. Therefore, we examined whether the humoral and/or cellular immune response after influenza vaccination differed between fatigued patients and non-fatigued individuals and between PCF and CFS patients. Methods: PCF (n=15) and CFS patients (n=22) were vaccinated against influenza. Age and gender matched non-fatigued cancer survivors (n=12) and healthy controls (n=23) were included for comparison. Antibody responses were measured at baseline and at day 21 by a hemagglutination inhibition test. T cell responses were measured at baseline and at day 7 by a lymphocyte proliferation and activation assay. Results: Both patient groups developed seroprotection rates comparable to the accompanying control groups. Functional T cell reactivity was observed in all groups. Proliferation at baseline was significantly lower in fatigued patients compared to non-fatigued individuals. A significant increase in proliferation from baseline to day 7 was observed in fatigued patients, but not in controls. At day 7, proliferation was not significantly different between fatigued patients and non-fatigued individuals. CD4+CD127-FoxP3+ expression was significantly higher in PCF patients compared to non-fatigued cancer survivors. Conclusions: We observed a lower T cell proliferation at baseline in fatigued patients compared to non-fatigued individuals, suggesting a difference in the baseline state of the immune system between fatigued patients and non-fatigued individuals. Furthermore, the difference in CD4+CD127-FoxP3+ expression between PCF and CFS patients suggests subtle differences in immune state between these two fatigued patient groups. However, since humoral and cellular immune responses after vaccination did not differ significantly between fatigued patients and non-fatigued individuals, vaccination of fatigued patients (PCF and CFS) can be effective.
APA, Harvard, Vancouver, ISO, and other styles
6

Schluederberg, Ann. "Chronic Fatigue Syndrome Research." Annals of Internal Medicine 117, no. 4 (August 15, 1992): 325. http://dx.doi.org/10.7326/0003-4819-117-4-325.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

EBARA, Ryuichiro. "Research on Corrosion Fatigue." Transactions of the Japan Society of Mechanical Engineers Series A 72, no. 720 (2006): 1119–22. http://dx.doi.org/10.1299/kikaia.72.1119.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

ZHANG, WeiYue, ChenXia HAN, SiYuan GUO, Feng LI, FengZhi WU, Xi YANG, and Jie MA. "Modern research of fatigue." SCIENTIA SINICA Vitae 46, no. 8 (August 1, 2016): 903–12. http://dx.doi.org/10.1360/n052016-00019.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Bower, Julienne E., Patricia A. Ganz, Katherine A. Desmond, Julia H. Rowland, Beth E. Meyerowitz, and Thomas R. Belin. "Fatigue in Breast Cancer Survivors: Occurrence, Correlates, and Impact on Quality of Life." Journal of Clinical Oncology 18, no. 4 (February 14, 2000): 743. http://dx.doi.org/10.1200/jco.2000.18.4.743.

Full text
Abstract:
PURPOSE: To describe the occurrence of fatigue in a large sample of breast cancer survivors relative to general population norms and to identify demographic, medical, and psychosocial characteristics of fatigued survivors. PATIENTS AND METHODS: Breast cancer survivors in two large metropolitan areas completed standardized questionnaires as part of a survey study, including the RAND 36-item Health Survey, Center for Epidemiological Studies–Depression Scale, Breast Cancer Prevention Trial Symptom Checklist, Medical Outcomes Study Sleep Scale, and demographic and treatment-related measures. RESULTS: On average, the level of fatigue reported by the breast cancer survivors surveyed (N = 1,957) was comparable to that of age-matched women in the general population, although the breast cancer survivors were somewhat more fatigued than a more demographically similar reference group. Approximately one third of the breast cancer survivors assessed reported more severe fatigue, which was associated with significantly higher levels of depression, pain, and sleep disturbance. In addition, fatigued women were more bothered by menopausal symptoms and were somewhat more likely to have received chemotherapy (with or without radiation therapy) than nonfatigued women. In multivariate analyses, depression and pain emerged as the strongest predictors of fatigue. CONCLUSION: Although the majority of breast cancer survivors in this large and diverse sample did not experience heightened levels of fatigue relative to women in the general population, there was a subgroup of survivors who did report more severe and persistent fatigue. We identified characteristics of these women that may be helpful in elucidating the mechanisms underlying fatigue in this population, as well as directing intervention efforts.
APA, Harvard, Vancouver, ISO, and other styles
10

de Raaf, Pleun J., Cora de Klerk, Reinier Timman, Jan J. V. Busschbach, Wendy H. Oldenmenger, and Carin C. D. van der Rijt. "Systematic Monitoring and Treatment of Physical Symptoms to Alleviate Fatigue in Patients With Advanced Cancer: A Randomized Controlled Trial." Journal of Clinical Oncology 31, no. 6 (February 20, 2013): 716–23. http://dx.doi.org/10.1200/jco.2012.44.4216.

Full text
Abstract:
Purpose Several guidelines on the treatment of cancer-related fatigue recommend optimizing treatment of accompanying symptoms. However, evidence for this recommendation from randomized clinical trials is lacking. We investigated whether monitoring and protocolized treatment of physical symptoms alleviates fatigue. Patients and Methods In all, 152 fatigued patients with advanced cancer were randomly assigned to protocolized patient-tailored treatment (PPT) of symptoms or care as usual. The PPT group had four appointments with a nurse who assessed nine symptoms on a 0 to 10 numeric rating scale (NRS). Patients received a nonpharmacologic intervention for symptoms with a score ≥ 1 and a medical intervention for symptoms with a score ≥ 4. Fatigue dimensions, fatigue NRS score, interference of fatigue with daily life, symptom burden, quality of life, anxiety, and depression were measured at baseline and after 1, 2, and 3 months. Differences between the groups over time were assessed by using mixed modeling. Results Seventy-six patients were randomly assigned to each study arm. Mean age was 58 ± 10 years, 57% were female, and 65% were given palliative chemotherapy. We found significant improvements over time in favor of PPT for the primary outcome general fatigue (P = .01), with significant group differences at month 1 (effect size, 0.26; P = .007) and month 2 (effect size, 0.35; P = .005). Improvements in favor of PPT were also found for the following secondary outcomes: fatigue dimensions “reduced activity” and “reduced motivation,” fatigue NRS, symptom burden, interference of fatigue with daily life, and anxiety (all P ≤ .03). Conclusion In fatigued patients with advanced cancer, nurse-led monitoring and protocolized treatment of physical symptoms is effective in alleviating fatigue.
APA, Harvard, Vancouver, ISO, and other styles
11

Chen, Binglin, Tianlai Yu, Linlin Zhang, Yuxuan Wu, and Yifan Wang. "Research on Fatigue Performance of the Rib Beam Bridge Carriageway Slab Based on Cumulative Damage Theory." Advances in Civil Engineering 2022 (December 5, 2022): 1–10. http://dx.doi.org/10.1155/2022/7455038.

Full text
Abstract:
The purpose of this paper is to study the development law of fatigue damage of the carriageway slab of reinforced concrete ribbed girder bridge and to provide theoretical support for carriageway slab design. The fatigue test of reinforced concrete rib beam bridge carriageway slabs was conducted. Based on the material fatigue damage theory, the fatigue damage model was established by ABAQUS software. The fatigue performance, failure mechanism, and the effect of fatigue level on the fatigue performance of the carriageway slab were studied. The experimental research results show that when the slab of the rib beam bridge was fatigued, and the radial crack appeared at the bottom of the slab, which belonged to punching shear failure. The fatigue failure process was divided into three stages as follows: in the initial stage of fatigue, the deflection and reinforcement strain increased linearly with fatigue times, and in the stable development stage, the bottom deflection and reinforcement strain increased steady progression. With the increase in fatigue cycles, the fatigue accumulation damage was gradually accumulated, and the stiffness of the carriageway slab gradually decreases. In the rapid development stage, the deflection, reinforcement strain, and crack growth speed were significantly increased. According to the fatigue damage model, the simulation analysis results show that the accumulated damage to the bridge carriageway was more serious with the increase in fatigue load level. The fatigue damage degree was lower when the design fatigue load level is not greater than 0.55.
APA, Harvard, Vancouver, ISO, and other styles
12

Tomasini, Pascale, Catherine Brown, Ashlee Vennettilli, Aein Zarrin, Aditi Dobriyal, Linda Chen, Maryam Mirshams, et al. "Assessing the utility of a patient-reported screener question to detect fatigue symptoms: Improving the quality of systematic symptom measurement in clinical practice." Journal of Clinical Oncology 32, no. 30_suppl (October 20, 2014): 236. http://dx.doi.org/10.1200/jco.2014.32.30_suppl.236.

Full text
Abstract:
236 Background: In Ontario, there is a concerted effort to screen all cancer outpatients for clinically significant symptoms at every visit, without causing undue burden on the patient. Although fatigue symptoms are common in cancer patients, severe fatigue may require clinical intervention. To reduce reporting fatigue, we evaluated the use of a single item screener question to detect severe fatigue, as defined through the FACT-Fatigue Scale (FACT-F), with the goal of reducing patient reporting burden. Methods: 316 Princess Margaret Cancer Centre outpatients across a wide range of cancers at all phases of therapies and disease stages were asked to report fatigue symptoms using the FACT-F. The ability of one screener question “I feel fatigued” to detect severe fatigue symptoms in any of the six other fatigue-related questions was evaluated. Using the presence of any severe fatigue symptom as the reference, sensitivity (Se) and specificity (Sp) of the screener question was determined. Results: Median age was 59 (19-91) years; 45% were male. The prevalence of significant, severe fatigue-related symptoms for the six individual questions covering various fatigue domains ranged from 4% to 7%. 12% of patients exhibited at least one severe symptom on FACT-F (prevalence). Defining a positive screen as “quite a bit” or “very much” fatigued, with 16% prevalence, the screener question was able to correctly identify any severe symptom 81% of the time (Se) and was able to rule out any severe symptoms 92% of the time (Sp). Conclusions: The use of a screener question to accurately detect patient symptoms provide patients with the ability to be involved in their care without being overly burdened in the process. At the meeting we will provide updated results on 500 patients, the potential modifying role of clinico-demographic factors, and results of the performance two additional screener questions on fatigue. While patient-reported outcomes are widely used in research, they may also be a practical and acceptable means to accurately detect clinically important symptoms in the clinic.
APA, Harvard, Vancouver, ISO, and other styles
13

Perry, Edward Belk, Sudhanshu Bharat Mulay, Jayesh Kamath, Robert J. Dowsett, Jacob Neuwirth, James Grady, Courtney Gold, Bruce Liang, and Susan Tannenbaum. "Factors influencing fatigue in breast cancer patients undergoing breast irradiation." Journal of Clinical Oncology 33, no. 28_suppl (October 1, 2015): 96. http://dx.doi.org/10.1200/jco.2015.33.28_suppl.96.

Full text
Abstract:
96 Background: We are conducting an exploratory, prospective study to investigate factors associated with radiation-induced fatigue in women with early breast cancer undergoing radiation therapy (RT) for breast conservation. Our hypothesis is that fatigue associated with adjuvant RT is related to tissue damage by apoptosis and inflammation, and to baseline psychological profiles. Methods: All subjects were assessed immediately before RT (T1), mid-point of RT (T2), end of RT (T3), 6 months (T4) and 1 year (T5) after completing RT. Clinical evaluations of skin toxicity, laboratory measures, fatigue, distress, depression, anxiety, sleep, energy and pain were assessed using validated measures. Results: Thirty-one subjects (target 50) have been enrolled to date; 23 have completed T3 and are presented here. 35% had a history of depression and 17% are currently depressed; 22% had past and current anxiety. BCTOS breast pain, tenderness and sensitivity increased during RT and were consistent with functional pain scores. No clinically relevant changes in energy or distress were seen. Despite a notable prevalence of depression and anxiety at baseline, there were no relevant changes in depression or anxiety. Most IL-1β, IL-4, IL-6 and IL-10 levels were undetectable. 39% of subjects showed increases in fatigue and were 14 years younger than non-fatigued subjects. Fatigued subjects had more frequent past (44%) and current (33%) depression than non-fatigued subjects (29% and 7%, respectively), and woke up tired more frequently (44% versus 0%). Current depression, but not fatigue, was associated with decreased a.m. cortisol levels during RT. 56% of fatigued subjects had an RTOG Acute Skin Toxicity score > 1 and 78% had BCTOS breast-specific pain subscales scores ≥ 3 during RT vs. 21% and 14% of non-fatigued subjects, respectively. Serum caspase-1 (inflammation) and caspase-3 (apoptosis), hs-CRP and TNF-α were increased but not associated with fatigue. Conclusions: Our preliminary findings suggest that radiation-associated fatigue is complex and, as hypothesized, is dependent on a patient’s psychological profile in the setting of skin toxicity and insomnia. If this data holds at study completion, it may guide clinical interventions.
APA, Harvard, Vancouver, ISO, and other styles
14

Liu, Fangping, and Jianting Zhou. "Research on Fatigue Strain and Fatigue Modulus of Concrete." Advances in Civil Engineering 2017 (2017): 1–7. http://dx.doi.org/10.1155/2017/6272906.

Full text
Abstract:
Concrete fatigue strain and fatigue modulus evolution play a vital role in the evaluation of the material properties. In this paper, by analyzing the advantages and disadvantages of existing concrete strain analysis methods, the level-S nonlinear fatigue strain model was proposed. The parameters’ physical meaning, the ranges, and the impact on the shape of the curve were all discussed. Then, the evolution model of fatigue modulus was established based on the fatigue strain evolution model and the hypothesis of fatigue modulus inversely related fatigue strain amplitude. The results indicate that the level-S model covered all types of fatigue strain evolution. It is very suitable for the description of strain evolution of concrete for its strong adaptability and high accuracy. It was found that the fitting curves coincided with the experimental curves very well, and the correlation coefficients were all above 0.98. The evolution curves of fatigue strain modulus both have three stages, namely, variation phase, linear change stage, and convergence stage. The difference is that the fatigue strain evolution curve is from the lower left corner to the upper right corner, but the fatigue modulus evolution curve is from the upper left corner to the right lower corner.
APA, Harvard, Vancouver, ISO, and other styles
15

Heras, P., A. Hatzopoulos, K. Kritikos, and S. Karagiannis. "Level of fatigue in colorectal cancer patients (ccp) receiving adjuvant chemotherapy." Journal of Clinical Oncology 27, no. 15_suppl (May 20, 2009): e20612-e20612. http://dx.doi.org/10.1200/jco.2009.27.15_suppl.e20612.

Full text
Abstract:
e20612 The aim of the study was to assess the level of fatigue during and after adjuvant chemotherapy in ccp. Methods: 52 patients (age 30–73 years) with colorectal cancer were recruited between May 2005 and November 2008. Fatigue intensity was measured according to 10-score visual analog scale (VAS, 0-no fatigue, 10-fatigue as bad as it could possibly be) before the start of adjuvant chemotherapy, once weekly during chemotherapy, 20 days and 6 months after. Results: Over half of the patients (56%) demonstrated no fatigue before the start of adjuvant chemotherapy. Fatigue intensity increased gradιιally during chemotherapy. It highest in the last week of treatment 20 days after the end of chemotherapy, fatigue intensity was still higher than before treatment, but 6 months after it was lower than the pretreatment level (mean fatigue-VAS was 1.17, 2.41, 1.45 and 0.61, respectively). However, 17% of the patients defined their fatigue as 2 or more in VAS six months after chemotherapy had ended. About 1/4 of the patients reported nο fatigue during chemotherapy. Increased level of fatigue was associated with an increased need of rest. Patients with pain (pain-VAS 1 or more) manifested higher fatigue level in comparison with individuals without pain (mean fatigue-VAS 3.20 vs 1.02 in the last week of treatment). Older patients (>65 years) estimated their fatique in the last week of treatment on lower level than younger patients (fatigue-VAS 1.79 vs. 2.83 No significant financial relationships to disclose.
APA, Harvard, Vancouver, ISO, and other styles
16

Nonaka, Isamu. "Some issues in creep-fatigue research." Strength, Fracture and Complexity 13, no. 4 (April 27, 2021): 163–75. http://dx.doi.org/10.3233/sfc-204004.

Full text
Abstract:
In the component operated at elevated temperatures, the life evaluation should be made in consideration of both creep and fatigue (creep-fatigue) such as the linear damage summation rule. However, the concept of creep-fatigue life evaluation has not spread well in the industry. In order to consider the reason, a series of past creep-fatigue research was surveyed, namely experimental methods, life evaluation procedures and strength design guidelines. As a result, it was revealed that the mechanism of creep-fatigue interaction has not been fully clarified yet, which results in obscuring the necessity of creep-fatigue life evaluation. The necessity of creep-fatigue life evaluation was reviewed and consequently it proved to be necessary in two cases. One is the case where the creep-fatigue interaction is significant for some kinds of material, loading modes and temperatures. The other is one where the amount of creep damage is almost the same as that of fatigue damage even though the creep-fatigue interaction is insignificant.
APA, Harvard, Vancouver, ISO, and other styles
17

Torres, Mylin Ann, Thaddeus Pace, Jennifer Felger, Tian Liu, Karen D. Godette, Liza Jane Stapleford, Donna Mister, and Andrew H. Miller. "A prospective longitudinal study of cancer-related fatigue in patients undergoing breast-conserving surgery and radiation with or without chemotherapy for breast cancer." Journal of Clinical Oncology 30, no. 15_suppl (May 20, 2012): 9122. http://dx.doi.org/10.1200/jco.2012.30.15_suppl.9122.

Full text
Abstract:
9122 Background: We prospectively evaluated risk factors for persistent cancer-related fatigue in women with breast cancer undergoing lumpectomy with or without chemotherapy (CTX) prior to whole breast radiotherapy (XRT). We assessed the potential role of inflammatory mediators, demographic characteristics, and treatment history including CTX. Methods: Following lumpectomy, 60 women received a definitive course of whole breast XRT (50 Gy plus a 10 Gy boost). Prior to XRT, at week 6 of XRT, and 6 weeks post XRT, subjects completed the Multidimensional Fatigue Inventory (MFI) and underwent blood draws for inflammatory mediators (protein and mRNA). Results: Independent multivariate analyses of clinical and demographic factors revealed that CTX (p<.001) , given neoadjuvantly or adjuvantly, and age <50 (p=.03) were significant predictors of higher fatigue scores post XRT. Mean MFI scores in patients treated with CTX (n=24) were 20 points higher than patients not treated with CTX (p<.001) with a clinically meaningful difference in scores being 10 points on the MFI. Gene ontology analysis of differentially expressed genes indicated increased activation of genes involved in immune and inflammatory responses in fatigued vs. non-fatigued patients (p<.001). Of the inflammatory mediators, plasma IL-6 prior to XRT was the strongest predictor of post XRT fatigue (p=.02). Moreover, plasma IL-6 concentrations prior to XRT were significantly higher in patients who received CTX (mean 4.96 vs. 2.53, p=.01). Patients who received CTX also had significantly higher levels of NF Kappa B DNA binding 6 weeks post XRT (p<.001), and transcription factor binding analysis revealed a greater representation of genes with the NF Kappa B DNA binding motif in fatigued vs. non-fatigued patients (p =.05). Conclusions: Collectively, these data suggest an interaction between CTX and XRT leading to inflammation and fatigue several weeks post XRT. This relationship was independent of whether CTX was given pre or post-operatively. Treatments targeting inflammation before XRT may reduce fatigue post therapy, particularly in patients previously treated with CTX.
APA, Harvard, Vancouver, ISO, and other styles
18

Antonov, A. Yu, Yu S. Mets, and P. I. Fedorenko. "Research of microsecond fatigue blasting." Mining Journal of Kryvyi Rih National University, no. 105 (2019): 99–103. http://dx.doi.org/10.31721/2306-5435-2019-1-105-99-103.

Full text
APA, Harvard, Vancouver, ISO, and other styles
19

MORITA, Tatsuro. "Recent Trends of Fatigue Research." Journal of the Society of Materials Science, Japan 66, no. 5 (2017): 372–79. http://dx.doi.org/10.2472/jsms.66.372.

Full text
APA, Harvard, Vancouver, ISO, and other styles
20

NAKAMURA, Takashi. "Recent Trends of Fatigue Research." Journal of the Society of Materials Science, Japan 66, no. 6 (2017): 435–41. http://dx.doi.org/10.2472/jsms.66.435.

Full text
APA, Harvard, Vancouver, ISO, and other styles
21

KIKUCHI, Shoichi, and Jun KOMOTORI. "Recent Trends of Fatigue Research." Journal of the Society of Materials Science, Japan 66, no. 7 (2017): 535–41. http://dx.doi.org/10.2472/jsms.66.535.

Full text
APA, Harvard, Vancouver, ISO, and other styles
22

NAKAI, Yoshikazu, and Daiki SHIOZAWA. "Recent Trends of Fatigue Research." Journal of the Society of Materials Science, Japan 66, no. 8 (2017): 621–26. http://dx.doi.org/10.2472/jsms.66.621.

Full text
APA, Harvard, Vancouver, ISO, and other styles
23

UEMATSU, Yoshihiko. "Recent Trends of Fatigue Research." Journal of the Society of Materials Science, Japan 66, no. 9 (2017): 688–94. http://dx.doi.org/10.2472/jsms.66.688.

Full text
APA, Harvard, Vancouver, ISO, and other styles
24

Corwin, Elizabeth J., Laura Cousino Klein, and Kristin Rickelman. "Predictors of Fatigue in Healthy Young Adults: Moderating Effects of Cigarette Smoking and Gender." Biological Research For Nursing 3, no. 4 (April 2002): 222–33. http://dx.doi.org/10.1177/109980040200300407.

Full text
Abstract:
Fatigue is a common complaint of patients seen in primary care. Factors that contribute to fatigue in a patient population include poor health status, psychological stress, poor nutrition, and pregnancy. Less well understood are factors that contribute to fatigue among healthy, nonpregnant individuals. Within the framework of the theory of unpleasant symptoms, 40 healthy young smoking and nonsmoking adults between the ages of 18 and 35 were evaluated to determine self-report level of fatigue and contributing physiological, psychological, and situational factors. Results indicate that while self-report of fatigue did not vary in this population based on gender, subjects who were moderate to heavy cigarette smokers were significantly more fatigued than were nonsmokers (F= 10.24, df = 1, 38, P < 0.01), with the effect being specific to male smokers. Self-report of fatigue did not correlate with body mass index, baseline inflammatory or immune status, or blood pressure. Positive psychological and situational predictors of fatigue included depression (r = 0.556, P < 0.001), state anxiety (r = 0.569, P < 0.001), sleep quality (r = –0.399, P < 0.05), and sleep quantity (r = –0.411, P < 0.05). These results suggest that psychological and situational factors are key contributors to fatigue in young adults and that smoking is a risk factor for fatigue in men.
APA, Harvard, Vancouver, ISO, and other styles
25

Meeske, Kathleen A., Stuart E. Siegel, Denise R. Globe, Wendy J. Mack, and Leslie Bernstein. "Prevalence and Correlates of Fatigue in Long-Term Survivors of Childhood Leukemia." Journal of Clinical Oncology 23, no. 24 (August 20, 2005): 5501–10. http://dx.doi.org/10.1200/jco.2005.03.210.

Full text
Abstract:
Purpose To estimate the prevalence of fatigue, identify the factors associated with fatigue, and to explore the relationship between fatigue and quality of life (QOL) in long-term survivors of childhood acute lymphoblastic leukemia (ALL). Methods One hundred sixty-one ALL survivors diagnosed at Childrens Hospital Los Angeles (Los Angeles, CA) before age 18 years and between January 1, 1975 and December 31, 1995, participated in a structured telephone interview. Participants were aged 18 to 41 years and off treatment for an average of 14 years. Four measures of fatigue, including the Revised–Piper Fatigue Scale, were used to assess fatigue; depression was assessed using the Center for Epidemiological Studies Depression Scale. Multivariate logistic regression models were developed to identify factors associated with fatigue and depression. Results Prevalence of fatigue (30%) fell within the general population normal limits. Fatigue and depression were highly correlated (Pearson r = 0.75). Fatigue was associated with marriage (OR = 0.11; 95% CI, 0.02 to 0.50), having children (OR = 5.80; 95% CI, 1.30 to 25.82), sleep disturbances (OR = 6.15; 95% CI, 2.33 to 16.22), pain (OR = 5.56; 95% CI, 2.13 to 14.48), obesity (OR = 3.80; 95% CI, 1.41 to 10.26), cognitive impairment (OR = 2.56; 95% CI, 1.02 to 6.38), and exercise-induced symptoms (OR = 2.98; 95% CI, 1.11 to 8.02). Four factors associated with fatigue were also associated with depression: sleep disturbances, pain, obesity, and cognitive impairment. Fatigue was inversely related to QOL. Conclusion Some survivors of childhood ALL experience fatigue many years after treatment. Fatigued survivors represent a high-risk subgroup as they report more depression and poorer QOL than nonfatigued survivors and their peers.
APA, Harvard, Vancouver, ISO, and other styles
26

Zhu, Hong Bing, Bo Xia, and Yao Zhao. "RC Beam Bridge’s Fatigue Cumulative Damage Rule Research." Advanced Materials Research 787 (September 2013): 829–32. http://dx.doi.org/10.4028/www.scientific.net/amr.787.829.

Full text
Abstract:
Fatigue damage is the RC beam bridge is facing a big problem, for the RC beam bridge fatigue tests and fatigue cumulative damage theory research is very meaningful. Summarizes the research achievements of the RC beam bridge fatigue test, from constant amplitude fatigue, luffing fatigue and stochastic fatigue, etc, are discussed in this paper. Analyses the existing linear, nonlinear and probability fatigue cumulative damage theory and its applicable conditions, advantages and disadvantages. RC fatigue tests were discussed and the problems that exist in the fatigue cumulative damage theory research.
APA, Harvard, Vancouver, ISO, and other styles
27

Alfano, Catherine M., Ikuyo Imayama, Marian L. Neuhouser, Janice K. Kiecolt-Glaser, Ashley Wilder Smith, Kathleen Meeske, Anne McTiernan, et al. "Fatigue, Inflammation, and ω-3 and ω-6 Fatty Acid Intake Among Breast Cancer Survivors." Journal of Clinical Oncology 30, no. 12 (April 20, 2012): 1280–87. http://dx.doi.org/10.1200/jco.2011.36.4109.

Full text
Abstract:
PurposeEvidence suggests that inflammation may drive fatigue in cancer survivors. Research in healthy populations has shown reduced inflammation with higher dietary intake of ω-3 polyunsaturated fatty acids (PUFAs), which could potentially reduce fatigue. This study investigated fatigue, inflammation, and intake of ω-3 and ω-6 PUFAs among breast cancer survivors.MethodsSix hundred thirty-three survivors (mean age, 56 years; stage I to IIIA) participating in the Health, Eating, Activity, and Lifestyle Study completed a food frequency/dietary supplement questionnaire and provided a blood sample assayed for C-reactive protein (CRP) and serum amyloid A (30 months after diagnosis) and completed the Piper Fatigue Scale and Short Form-36 (SF-36) vitality scale (39 months after diagnosis). Analysis of covariance and logistic regression models tested relationships between inflammation and fatigue, inflammation and ω-3 and ω-6 PUFA intake, and PUFA intake and fatigue, controlling for three incremental levels of confounders. Fatigue was analyzed continuously (Piper scales) and dichotomously (SF-36 vitality ≤ 50).ResultsBehavioral (P = .003) and sensory (P = .001) fatigue scale scores were higher by increasing CRP tertile; relationships were attenuated after adjustment for medication use and comorbidity. Survivors with high CRP had 1.8 times greater odds of fatigue after full adjustment (P < .05). Higher intake of ω-6 relative to ω-3 PUFAs was associated with greater CRP (P = .01 after full adjustment) and greater odds of fatigue (odds ratio, 2.6 for the highest v lowest intake; P < .05).ConclusionResults link higher intake of ω-3 PUFAs, decreased inflammation, and decreased physical aspects of fatigue. Future studies should test whether ω-3 supplementation may reduce fatigue among significantly fatigued breast cancer survivors.
APA, Harvard, Vancouver, ISO, and other styles
28

Balducci, L., A. Luciani, M. Extermann, A. Cantor, and P. Jacobsen. "Fatigue is a cause of functional dependence in older cancer survivors." Journal of Clinical Oncology 24, no. 18_suppl (June 20, 2006): 8538. http://dx.doi.org/10.1200/jco.2006.24.18_suppl.8538.

Full text
Abstract:
8538 Background: The incidence and prevalence of cancer increase with age, but the impact of cancer and its treatment on the function of older cancer survivors is unknown. The aim of the study was to establish the prevalence of fatigue in older cancer patients off chemotherapy and the correlation of fatigue with Instrumental Activities of Daily Living (IADL) and Activities of Daily Living (ADL). Methods: We reviewed the cases of 214 individuals aged 70–90 with different cancers, subsequently processed through the Senior Adult Oncology Program (SAOP) over 4 months. All patients of the SAOP are evaluated for function, depression, comorbidity and fatigue, and all of them have a complete blood count. Depression is assessed with the geriatric depression scale (GDS); comorbidity with the cumulative illness rating scale for geriatrics (CIRS-G), and fatigue with the Fatigue symptom inventory (FSI). Results: We found that 82% of the patients reported some degree of fatigue and in 74% of cases fatigue interfered with their daily activities. The average fatigue severity was 5.38 ± 2.59; the FSI interference score 19.1 ± 16.1, the average number of days fatigued in a week 4.07 ± 2.6. Fatigue severity was positively correlated with CIRS_G score (p = 0.04), creatinine clearance (p = 0.02) performance status (PS) (p < 0.0001), GDS (p = 0.04); ADL (0.02); Minimental status (MMS) (p = 0.05); fatigue interference with Cr Clp = ).005); PD (p = 0.001) ADL (p = 0.02); GDS (p = 0.01), and fatigue frequency with CIRS-G (p = 0.04); Cr CL (p = 0.01), PS (p = 0.0001), ADL (p = 0.03); IADL (p = 0.03), and GDS(0.09). ADL dependence was related to intensity (0.02), interference (0.0001) and days of fatigue; IADL dependence to interference score (p = 0.0001) and days of fatigue (p = 0.017) and GDS to intensity (p = 0.004); interference score (p = 0. 006) and frequency of fatigue. Conclusions: Fatigue is common in long term older cancer survivors, and is associated with functional dependence and depression. Control of fatigue may improve the function of older cancer survivors No significant financial relationships to disclose.
APA, Harvard, Vancouver, ISO, and other styles
29

He, Xiao Jun, Jing Liu, Zhen Di Yi, and Yuan Quan Yang. "Development of Detection Research on Fatigue Driving." Applied Mechanics and Materials 641-642 (September 2014): 813–17. http://dx.doi.org/10.4028/www.scientific.net/amm.641-642.813.

Full text
Abstract:
This paper presents the current most common fatigue-driving detection methods. The advantages and disadvantages of these detection methods are compared with. Moreover, several major products of the current fatigue detection are listed briefly. Furthermore, the development trends of driving-fatigue detection technology are prospected. The author believes that driver fatigue testing standards need to be further clarified and the non-contact detection method of driving-fatigue needs to be developed deeply. Information fusion is an important orientation for driving fatigue and we should design the cost-efficient detection products for fatigue-driving.
APA, Harvard, Vancouver, ISO, and other styles
30

Donnelly, Elizabeth A., Paul Bradford, Matthew Davis, Cathie Hedges, Doug Socha, and Peter Morassutti. "Fatigue and Safety in Paramedicine." CJEM 21, no. 6 (August 13, 2019): 762–65. http://dx.doi.org/10.1017/cem.2019.380.

Full text
Abstract:
ABSTRACTObjectivesExtant research has established an empirical relationship between fatigue and safety-related outcomes. It is not clear if these findings are relevant to Canadian paramedicine. The purpose of this study was to determine if fatigue and shiftwork variables were related to safety outcomes in Canadian paramedics.MethodsA survey was conducted with ten paramedic services in Ontario with a 40.5% response rate (n = 717). Respondents reported levels of fatigue, safety outcomes (injury, safety compromising behaviours, and medical errors/adverse events), work patterns (types of shifts, hours worked weekly) and demographic characteristics. Univariate and logistic regression analyses were used to assess for significant differences.ResultsIn this sample, 55% of paramedics reported being fatigued at work. Fatigued paramedics were over twice as likely to report injuries, three times as likely to report safety compromising behaviors, and 1.5 times more likely to report errors/adverse outcomes. When controlling for fatigue, shift length variables did not consistently influence safety outcomes.ConclusionThese results create preliminary evidence of a relationship between fatigue and safety outcomes in Canadian paramedicine. While more research is needed, these findings point to the influence fatigue has on safety outcomes and provide an indication that fatigue mitigation efforts may be worthwhile.
APA, Harvard, Vancouver, ISO, and other styles
31

Steingräber, Maria, and Petra Feyer. "Tumorbedingte Fatigue." Deutsche Zeitschrift für Onkologie 37, no. 02 (June 2005): 52–57. http://dx.doi.org/10.1055/s-2005-862561.

Full text
APA, Harvard, Vancouver, ISO, and other styles
32

Loge, Jon Håvard, Arne Foss Abrahamsen, Øivind Ekeberg, and Stein Kaasa. "Hodgkin's Disease Survivors More Fatigued Than the General Population." Journal of Clinical Oncology 17, no. 1 (January 1999): 253. http://dx.doi.org/10.1200/jco.1999.17.1.253.

Full text
Abstract:
PURPOSE: To estimate the level of fatigue and frequency of fatigue cases among Hodgkin's disease survivors (HDS) and compare them with normative data from the general population. PATIENTS AND METHODS: A cross-sectional follow-up study was done of 557 HDS (age range, 19 to 74 years) treated at the Norwegian Radium Hospital from 1971 to 1991. The sample was approached by mail, and their data were compared with normative data from 2,214 controls (age range, 19 to 74 years) representative of the general Norwegian population. Of the 557 HDS, 459 (82%) responded. The mean age (± SD) at the time of study was 44 ± 12 years, and the mean observation time was 12 ± 6 years. The Fatigue Questionnaire (11 items) measures physical and mental fatigue. Two systems of scoring were used, dichotomized (0, 0, 1, and 1) and Likert (0, 1, 2, and 3). Total fatigue (TF) constitutes the sum of all the Likert scores. Caseness was defined as a total dichotomized score of ≥4 and fatigue that lasted 6 months or longer. RESULTS: The HDS had significantly higher levels of TF than the controls (14.3 v 12.2) (P < .001). Fatigue among the HDS equaled that of the controls in poorest health. More HDS (61%) than controls (31%) reported fatigue symptoms lasting 6 months or longer (P < .001). Fatigue cases were more frequent among HDS (men, 24%; women, 27%) than among the controls (men, 9%; women, 12%) (P < .001). Disease stage/substage IB/IIB predicted fatigue caseness (P = .03). No significant associations were found between treatment characteristics and fatigue. CONCLUSION: Hodgkin's disease survivors are considerably more fatigued than the general population and report fatigue of a substantially longer duration.
APA, Harvard, Vancouver, ISO, and other styles
33

Soriano, Maria, Alejandro Alvarez-Bustos, Javier Ramos, Carmen Piquin, Pablo Osorio, Javier Ros, Miguel Ramírez, et al. "Cancer-related fatigue in cancer survivors (CS) with no active treatment." Journal of Clinical Oncology 35, no. 31_suppl (November 1, 2017): 238. http://dx.doi.org/10.1200/jco.2017.35.31_suppl.238.

Full text
Abstract:
238 Background: Fatigue is a subjective experience that should be systematically assessed at the initial visit, at regular intervals and as clinically indicated. Quality of fatigue management should be included in institutional continuous quality improvement projects. The PERFORM questionnaire (PQ) was developed among Spanish speaking patients for the assessment of fatigue. Methods: Outpatients recruited for projects in which physical condition and physical activity (PA) were evaluated, rated their fatigue severity on the PQ (12-60, being 60 no fatigue). PHUEM-01 evaluated early colon CS after finishing adjuvant treatment (AT). PHUEM-02 early breast CS at the end of AT. PH-UEM-03 evaluated colorectal CS (localized and metastatic) at the time of diagnosis. MS-04 evaluated fatigue at the time of diagnosis of different tumors (localized and metastatic). Physical condition was evaluated through the one-mile walk test (VO2MAX) and handgrip dynamometer. Heart rate was measured as a sign of autonomic dysfunction. PA was objectively evaluated through accelerometers generating weekly MVPA information. Results: 262 CS (63% women) were recruited, (110 breast, 119 colorectal, 14 other), mean age 60. 30% of survivors reported no fatigue. Mean PERFORM score 48,34; VO2max 25.5 ml/kg/min; handgrip 29,14 kg; HR 75 bpm. Women, younger CS, CS with worse physical condition, higher BMI, higher HR and less active CS reported more fatigue (only sex, handgrip strength and weekly PA were statistically significant for the overall population). Breast CS reported higher scores of fatigue than colorectal CS (p < 0.000). For breast CS there was an association between fatigue and HR which was not seen in the overall population. Metastatic colorectal did not report worse scores of fatigue than CS with localized tumors. Conclusions: As a subjective symptom, fatigue is unpredictable and its severity and interference with daily activities can be bothersome in an unexpected population (young CS, early stages) with no active treatment. More physically active and fitter CS are less fatigued. Mechanisms underlying cancer-related fatigue could be different for different tumors. More research in prevalence, evaluation, mechanisms and management of fatigue is needed.
APA, Harvard, Vancouver, ISO, and other styles
34

Knobel, Heidi, Jon Håvard Loge, May Brit Lund, Kolbjørn Forfang, Ole Nome, and Stein Kaasa. "Late Medical Complications and Fatigue in Hodgkin’s Disease Survivors." Journal of Clinical Oncology 19, no. 13 (July 1, 2001): 3226–33. http://dx.doi.org/10.1200/jco.2001.19.13.3226.

Full text
Abstract:
PURPOSE: Long-term medical complications, such as cardiac, pulmonary, and thyroid dysfunction, are frequent among Hodgkin’s disease survivors (HDSs). Chronic fatigue is also highly prevalent among HDSs. Few studies have explored possible etiologic explanations for fatigue. The aim of this study was to explore whether late cardiac, pulmonary, and thyroid complications after curative treatment for Hodgkin’s disease (HD) may explain the high level of fatigue among HDSs. PATIENTS AND METHODS: Four-hundred fifty-nine patients treated for HD at the Norwegian Radium Hospital from 1971 to 1991 were included in a cross-sectional, follow-up study of subjective health status. Fatigue (physical [PF] and mental), was measured by the Fatigue Questionnaire. A subcohort of the HDSs (116 patients) treated from 1980 to 1988 were included in a separate study in which long-term cardiac, pulmonary, and thyroid complications were assessed. All patients had received radiotherapy, and 63 patients had received additional chemotherapy. The present study comprised 92 patients (mean age, 37 years; range, 23 to 56 years) who participated in both studies. RESULTS: HDSs with pulmonary dysfunction were more fatigued than HDSs with normal pulmonary function (PF 10.9 v 8.9; P < .05). Gas transfer impairment was the most prevalent pulmonary dysfunction, and three times as many patients with gas transfer impairment reported chronic fatigue (duration, 6 months or longer), compared with patients without pulmonary dysfunction (48% v 17%, P < .01). No associations were found between cardiac sequelae or hypothyroidism and fatigue. CONCLUSION: Pulmonary dysfunction is associated with fatigue in HDSs. Cardiac sequelae was not associated with fatigue in HDSs. We question the absence of an association between thyroid complications and fatigue.
APA, Harvard, Vancouver, ISO, and other styles
35

Hackney, Alisha J., N. Jennifer Klinedinst, Barbara Resnick, and Meg Johantgen. "Association of Systemic Inflammation and Fatigue in Osteoarthritis: 2007−2010 National Health and Nutrition Examination Survey." Biological Research For Nursing 21, no. 5 (June 25, 2019): 532–43. http://dx.doi.org/10.1177/1099800419859091.

Full text
Abstract:
Background/Purpose:Mechanistic insight into osteoarthritis fatigue is needed as clinical management of this condition is nonspecific. Systemic inflammation is associated with fatigue in other chronic diseases. The purpose of this study was to explore the relationship between systemic inflammation and fatigue in osteoarthritis, while controlling for covariates.Method:This secondary analysis with a cross-sectional, multiyear retrospective design used data from the 2007−2010 National Health and Nutrition Examination Survey. Adults with self-reported osteoarthritis who participated in an examination at a mobile center and had no comorbidities associated with fatigue or systemic inflammation were included ( n = 296). Complex sample analysis, independent samples t tests, and χ2tests of independence were used to explore differences between nonfatigued and fatigued adults with osteoarthritis. Adjusted hierarchical logistic regression models were used to calculate odds of fatigue as a function of two systemic inflammatory markers, C-reactive protein (CRP), and white blood cell (WBC) count.Results:Fatigued adults with osteoarthritis had significantly higher CRP levels and WBC counts compared to nonfatigued adults with osteoarthritis. In adjusted logistic regression models, increased CRP was associated with higher odds of fatigue when controlling for age, days affected by pain, depressive symptoms, sleep quantity, and body mass index (Odds ratio [ OR] = 3.38, 95% CI [1.18, 9.69]). WBC count was not associated with higher odds of fatigue when controlling for these variables ( OR = 1.10, 95% CI [0.92, 1.32]).Conclusion:Systemic inflammation may have a relationship with fatigue in osteoarthritis. Future work is necessary to replicate these findings in more robust studies.
APA, Harvard, Vancouver, ISO, and other styles
36

Matias, Margarida, Giulia Baciarello, Mohamed Neji, Stefan Michiels, Ann H. Partridge, Marc Karim Bendiane, Karim Fizazi, Michel Ducreux, Fabrice Andre, and Ines Maria Vaz Duarte Luis. "Fatigue and health behaviors in cancer survivors: A cross-sectional population based study." Journal of Clinical Oncology 35, no. 15_suppl (May 20, 2017): 10069. http://dx.doi.org/10.1200/jco.2017.35.15_suppl.10069.

Full text
Abstract:
10069 Background: A substantial proportion of breast, colo-rectal and prostate cancer patients (pts) can expect long term disease free survival after their primary treatment. Among those, fatigue commonly persists after diagnosis (dx) and can be debilitating. In this study, we evaluated the incidence of fatigue 2 years (y) after cancer dx and its association with health behaviors. Methods: We used a French population based cross-sectional study, which included a representative sample of pts from 12 cancer types (VICAN2). VICAN2 surveyed 4347 pts 2 y after dx. There is a 99% completion rate of fatigue related questions. For this study, we included 2017 pts with breast (1237), colo-rectal (348) and prostate (432) cancer without evidence of metastases at dx or relapse 2y after dx and with fatigue information. Severe fatigue was defined as average score of EORTC QLQC30 fatigue subscale ≥40 at 2y after dx. There were <1% of missing values in the evaluated covariates. Multivariate logistic regression models looked at associations of fatigue with heatlh behaviors (Δ in exercise since dx, exercise at diagnosis, body mass index (BMI), Δ since dx, smoking), age, gender, comorbidities, education, employment, cancer type, radiation, chemotherapy, hormonal therapy. Results: 52% of pts reported severe fatigue at 2 y after dx (median fatique score: 44, range: 0-100). 2 y after dx, 47% of pts either stopped or decrease exercise and 16% had a ≥ 10% change in BMI. Factors associated with fatigue included health behaviors (Table), but also age (adjusted odds ratio [ aOR] for severe fatigue, 95% confidence interval [95% CI]: 0.97, 0.96-0.98), gender ( aOR, 95 CI male vs. female: 0.5, 0.3-0.8), comorbidities ( aOR, 95 CI yes vs. no: 2.0, 1.6-2.4) and treatment type ( aOR, 95 CI radiation vs. no: 1.5, 1.1-2.0). Conclusions: Fatigue continues to be a substantial problem for cancer survivors 2 y after dx. Some factors that may contribute to persistent fatigue (health behaviors) may be amenable to interventions. [Table: see text]
APA, Harvard, Vancouver, ISO, and other styles
37

Wang, Fuwang, Xiaogang Kang, Rongrong Fu, and Bin Lu. "Research on driving fatigue detection based on basic scale entropy and MVAR-PSI." Biomedical Physics & Engineering Express 8, no. 5 (July 5, 2022): 055005. http://dx.doi.org/10.1088/2057-1976/ac79ce.

Full text
Abstract:
Abstract In long-term continuous driving, driving fatigue is the main cause of traffic accidents. Therefore, accurate and rapid detection of driver mental fatigue is of great significance to traffic safety. In our study, the electroencephalogram (EEG) signals of subjects were preprocessed to remove interference signals. The Butterworth band-pass filter is used to extract the EEG signals of α and β rhythms, and then the basic scale entropy of α and β rhythms is used as driving fatigue characteristics. In addition, combined with the fast multiple autoregressive (MVAR) model and phase slope index (PSI), short-term data is used to accurately estimate the effective connectivity of EEG signals between different channels, and analyzed the causality flow direction in the left and right prefrontal regions of drivers at different driving stages. Further comprehensive analysis of the driver’s driving fatigue state in the continuous driving phase. Finally, the correlation coefficient value between the parameter pairs (basic scale entropy, clustering coefficient, global efficiency) is calculated. The results showed that the causality flow outflow degree of prefrontal lobe decreased during the transition from sober driving state to tired driving state. The left and right prefrontal lobes were the source of causality in sober driving state, and gradually became the target of causality with the occurrence of driving fatigue. The results showed that when transitioning from a waking state to a fatigued driving state, the causal flow direction out-degree value of the prefrontal cortex on a declining curve, and the left and right prefrontal cortex exhibited the causal source in the awake driving state, which gradually changed into the causal target along with the occurrence of driving fatigue. The three parameters of basic scale entropy, clustering coefficient and global efficiency are used as driving fatigue characteristics, and every two parameters have strong correlation. It shows that the combination of basic scale entropy and MVAR-PSI method can effectively detect the driver’s long-term driving fatigue state in continuous driving mode.
APA, Harvard, Vancouver, ISO, and other styles
38

Wang, Chunsheng, Lan Duan, Musai Zhai, Yuxiao Zhang, and Shichao Wang. "Steel bridge long-term performance research technology framework and research progress." Advances in Structural Engineering 20, no. 1 (July 28, 2016): 51–68. http://dx.doi.org/10.1177/1369433216646005.

Full text
Abstract:
To ensure structural sustainability, it is necessary to conduct steel bridge long-term performance study, including bridge design, evaluation, maintenance, and reinforcement technology. The research on steel bridge long-term performance is introduced in four aspects: (1) fatigue performance experimental study for full-scale orthotropic steel bridge decks in laboratory to study its fatigue failure mechanism, in order to improve fatigue design methodology and find rational reinforcement and maintenance method; (2) conducting steel bridge out-of-plane distortion-induced fatigue performance study, and developing cold reinforcement method; (3) performance study for base material and typical joint under long-term vehicle and environmental effect in aging steel bridges, and safety assessment and maintenance of existing steel bridge; (4) temperature gradient monitoring for steel box girder model to build the temperature design mode. Meantime, in-situ tests and monitoring are conducted for steel bridge long-term performance detection, assessment, and maintenance. The study results in this article build the research framework of steel bridge long-term performance preliminarily, which is the basis for steel bridge sustainable design, maintenance, and cold reinforcement methodology system.
APA, Harvard, Vancouver, ISO, and other styles
39

G Rooney, Alasdair, William Hewins, Amie Walker, Lisa Withington, Mairi Mackinnon, Sara Robson, Aimee Green, et al. "INNV-23. LIFESTYLE COACHING IS FEASIBLE AND IMPROVES PILOT OUTCOMES IN FATIGUED BRAIN TUMOUR PATIENTS: THE BT-LIFE (BRAIN TUMOURS, LIFESTYLE INTERVENTIONS, AND FATIGUE EVALUATION) MULTI-CENTRE, PHASE II RCT." Neuro-Oncology 22, Supplement_2 (November 2020): ii121. http://dx.doi.org/10.1093/neuonc/noaa215.506.

Full text
Abstract:
Abstract BACKGROUND Fatigue is common and disabling for brain tumour patients. We studied the feasibility of two innovative lifestyle coaching interventions for high fatigue. METHODS Multi-centre phase II feasibility RCT (ISRCTN17883425). Adult primary brain tumour outpatients reporting significant fatigue (Brief Fatigue Inventory [BFI] score 4+), were randomised to one of three arms: Control; Health Coaching (“HC”, comprising eight structured coaching sessions on lifestyle behaviours); or HC plus Activation Coaching (“HC+AC”, adding two structured interviews targeting motivation to change). Outcomes were measured at baseline (T0), after interventions (T1), and at 16 weeks (T2). The primary outcome of feasibility was required for both recruitment (aim: average n= 5 fatigued patients recruited/month) and retention (aim: minimum 60% retention at T2). Secondary pilot outcomes included change in fatigue, depressive symptom, and QOL measures. RESULTS Over a nine-month recruitment period, n= 46 fatigued brain tumour patients were recruited (average n=5.1/month) and n= 34 were retained to endpoint (retention at T2= 73%), meeting the primary outcome of feasibility. Surprisingly, fatigue reduced significantly after HC (T1 mean change in BFI score from T0 baseline, relative to the equivalent change in control group: HC= -2.3 points [95%CI -3.4/-0.3]; HC+AC= -2.0 [-2.9/+0.1]; ANOVA p= 0.02) and was reduced in magnitude in both intervention groups at T2 (p= N.S). Both interventions also improved depressive symptoms (T1 mean change in HADS-Depression: HC= -2.0 points [-5.6/-0.1]; HC+AC= -2.9 [-6.5/-1.0]; Kruskal-Wallis p= 0.02). Patient-nominated QOL outcomes improved persistently after HC+AC (T2 mean change in PSYCHLOPS score: HC= -2.4 points [-5.4/+2.8]; HC+AC= -6.1 [-9.2/-0.8]; ANOVA p= 0.01). CONCLUSION Innovative coaching interventions, focused on lifestyle factors, are feasible to deliver to fatigued brain tumour patients. Preliminary signals suggest that these non-drug approaches may benefit several mediators of quality of life and warrant further study.
APA, Harvard, Vancouver, ISO, and other styles
40

Zhang, Dan Feng, Xiao Ming Tan, Jia Rui Qi, and Yan Li Li. "Research on Fatigue Notch Factor of Aluminum Corrosion Pits." Applied Mechanics and Materials 633-634 (September 2014): 141–44. http://dx.doi.org/10.4028/www.scientific.net/amm.633-634.141.

Full text
Abstract:
In the midst of highly corrosive marine environment the corrosion damage is very critical in the aircraft structure. Corrosion will reduce the fatigue strength and fatigue life of aircraft structure. The fatigue notch factor is a very important factor, and which directly affects the accuracy of the fatigue life estimation result. The LY12CZ aluminum alloy of aircraft structure was accelerated tested according to equivalent accelerated corrosion testing spectrum based on key environment data, then made a fatigue test. According to the fatigue test data fatigue notch factor was calculated and then the variation law of fatigue notch factor with the corrosion life was studied.
APA, Harvard, Vancouver, ISO, and other styles
41

Wang, Tie, Hong Mei Li, Rui Liang Zhang, and Zhi Fei Wu. "Research on Gear Contact Fatigue Stress Test." Applied Mechanics and Materials 86 (August 2011): 825–28. http://dx.doi.org/10.4028/www.scientific.net/amm.86.825.

Full text
Abstract:
This paper put forward the rapid measure method of the gear contact fatigue stress value with a few gear samples, which can get the estimated value of the gear fatigue limit value precisely and rapidly. And the gear fatigue life curve and fatigue damage accumulation curve are simulated by MATLAB. Comparing with the traditional test method, this method can reduce the cost and save the time.
APA, Harvard, Vancouver, ISO, and other styles
42

Mertens, Wilson C., Gregory Scott Emmons, Deborah Katz, Ruth Barham, Lucinda Jane Cassells, Saurabh Dahiya, and Leslie M. Howard. "Cancer cachexia syndrome (CS): A symptom complex associated with yet distinct from fatigue in ambulatory cancer patients." Journal of Clinical Oncology 30, no. 15_suppl (May 20, 2012): e19543-e19543. http://dx.doi.org/10.1200/jco.2012.30.15_suppl.e19543.

Full text
Abstract:
e19543 Background: Fatigue and CS (Lancet Oncol 2011;12:489) are common, multifactorial symptoms in ambulatory cancer patients. Clinical and laboratory correlates require further evaluation, although erythrocyte sedimentation rate (ESR), C-reactive protein (CRP), nausea and weight loss have been associated with both CS and fatigue. A practical classification of cancer-associated fatigue is lacking. Methods: Patients with cancer—newly diagnosed, receiving treatment, as well as those undergoing post-treatment surveillance and recurrent/persistent disease—were consented and screened with a visual aid. Those reporting moderate fatigue were asked to participate in a study assessing the degree and impact of fatigue using the FACT-G and study-specific questionnaires, followed by focused laboratory testing. Results: 102 patients were accrued. Mean age 60.2 years (95%CI: 58.0, 62.5); 80% female; no evidence of disease 55.9%; disease present/initial presentation 18.6%, persistent/recurrent disease 25.5%; >5% weight loss 35%; CS 39%; anorexia-CS 18%; increased ESR 49% and CRP 32%; currently receiving chemotherapy 16%. Patients currently receiving chemotherapy had more fatigue (p=.011); none of CS, >5% weight loss, disease presence/recurrence, histologic diagnosis, gender, age, pain, appetite, nausea, depression, exercise, blood pressure (BP), hemoglobin level, ESR or CRP were statistically associated with fatigue. CS was associated with increased ESR and reduced BP and appetite (all p<.001), and increased nausea and CRP (both p<.05) as well as disease presence/recurrence (p=.003). FACT-G total and physical scale scores did not differ between CS and other fatigued patients. Conclusions: CS affects a substantial proportion of ambulatory cancer patients screened for moderate fatigue; however, CS patients did not exhibit increased fatigue or poorer FACT-G scores despite differences in weight, disease extent, weight loss, or laboratory markers of inflammation when compared to other sampled patients. CS patients may represent a distinct subset of cancer fatigue, or this symptom complex may be distinct from fatigue even though present in the same patient.
APA, Harvard, Vancouver, ISO, and other styles
43

Zhou, Hong Yu, Yi Bo Chen, Jun Chang Ci, and Cong Kun Yang. "Research Status of Fatigue Damage Mechanism of Reinforced Concrete Beam." Applied Mechanics and Materials 858 (November 2016): 44–49. http://dx.doi.org/10.4028/www.scientific.net/amm.858.44.

Full text
Abstract:
Based on the fatigue damage mechanism, fatigue life, stiffness degradation, crack width change, bending, shear fatigue properties and other aspects, this paper introduces the research progress of the fatigue properties of ordinary reinforced concrete beams. And the existing reinforced concrete beam flexural, shear fatigue properties of research ideas, methods and results are summarized, providing the basis for further study on the fatigue performance of reinforced concrete beams. At present, the research results show that the fatigue damage of reinforced concrete beam is basically in accordance with the law of the three stages. In the early stage of fatigue, the tensile concrete cracks and exits, and the damage develops rapidly. In the middle of fatigue crack growth, fatigue damage is developed into a more moderate linear growth. In the late stage of fatigue, fatigue fracture occurs in the steel bar, and the bearing capacity of the beam is quickly lost.
APA, Harvard, Vancouver, ISO, and other styles
44

Mock, Victoria. "Fatigue management." Cancer 92, S6 (2001): 1699–707. http://dx.doi.org/10.1002/1097-0142(20010915)92:6+<1699::aid-cncr1500>3.0.co;2-9.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Park, Mi Sook, and Young Soon Byun. "An Analysis of Research on Fatigue." Journal of Nurses Academic Society 26, no. 4 (1996): 868. http://dx.doi.org/10.4040/jnas.1996.26.4.868.

Full text
APA, Harvard, Vancouver, ISO, and other styles
46

Palmer, Stephen. "Questionnaire fatigue and other research concerns." Counselling Psychology Review 14, no. 1 (February 1999): 31–34. http://dx.doi.org/10.53841/bpscpr.1999.14.1.31.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

Koevoets, Emmie, Sanne B. Schagen, Michiel B. de Ruiter, Mirjam I. Geerlings, Lenja Witlox, Elsken Van Der Wall, Martijn M. Stuiver, et al. "Effect of physical exercise on cognitive function after chemotherapy in patients with breast cancer: A randomized controlled trial (PAM study)." Journal of Clinical Oncology 39, no. 15_suppl (May 20, 2021): 12015. http://dx.doi.org/10.1200/jco.2021.39.15_suppl.12015.

Full text
Abstract:
12015 Background: Chemotherapy is associated with cognitive problems. Physical exercise is a promising intervention. We investigated whether exercise improves cognition in chemotherapy-exposed breast cancer (BC) patients 2-4 years after diagnosis. Methods: In the PAM study, we randomized chemotherapy-exposed BC patients with self-reported and test-confirmed cognitive problems to an exercise or control group. The 6-month exercise intervention consisted of 2 hours of supervised aerobic and resistance training and two hours of Nordic/power walking. Memory function measured with the Hopkins Verbal Learning Test-Revised (HVLT-R) was our primary outcome. Further measurements included online neuropsychological tests (Amsterdam Cognition Scan; ACS), self-reported cognitive complaints (MDASI-MM, EORTC QLQ C-30 cognitive functioning), physical fitness (VO2peak), fatigue (MFI, EORTC fatigue), quality of life (QoL; EORTC), anxiety (HADS) and depression (HADS, PHQ9). HVLT-R total recall was analyzed with a Fisher exact test for clinically relevant improvement of ≥5 words. Other outcomes were analyzed using multiple regression analyses adjusted for baseline and stratification factors. An hypothesis driven but not pre-specified analysis in patients with high baseline EORTC fatigue levels (≥39) was performed. Results: We randomized 181 patients to the exercise (n = 91) or control group (n = 90). Two-third of the patients attended ≥ 80% of the exercise program and physical fitness significantly improved compared to the control patients ( VO2peak1.4 ml/min/kg, 95% CI 0.6; 2.2). No difference in favor of the intervention group was seen on the primary cognitive outcome or other cognitive tests. However, significant beneficial intervention effects were found for self-reported cognition (MDASI-MM Severity (-0.7, -1.2;-0.1)), fatigue (general fatigue (-2.2, -3.3; -1.1), physical fatigue (-3.3, -4.4; -2.2), mental fatigue (-1.0, -2.0; 0.0), reduced motivation (-1.1, -2.0; -0.2) and reduced activity (-2.1, -3.2; -1.1)), QoL (summary score (4.0, 1.2; 6.7), global health status (5.8, 1.1; 10.6), role functioning (7.2, 1.3; 13.1) and social functioning (5.9, 0.2; 11.6)) and depression (PHQ9 (-1.16, -2.19; -0.13)). In high-fatigued patients, exercise did show significant positive effects on objective cognitive function (ACS Reaction Time (-26.8, -52.9; -0.6) and ACS Wordlist Learning (4.4, 0.5; 8.3)). Conclusions: A 6-month exercise intervention did not improve objectively measured cognitive function in chemotherapy-exposed BC patients with cognitive problems. However, self-reported cognitive function, physical fitness, fatigue, QoL and depression did improve. Unplanned analysis indicated a small positive effect of exercise on cognitive functioning in high-fatigued patients. Clinical trial information: NTR6104.
APA, Harvard, Vancouver, ISO, and other styles
48

Yu, Zhe Fu. "Research on the Fatigue Damage of the Asphalt Pavement." Applied Mechanics and Materials 256-259 (December 2012): 1776–79. http://dx.doi.org/10.4028/www.scientific.net/amm.256-259.1776.

Full text
Abstract:
Abstract: Fatigue cracking is a main form of structural damage of asphalt pavements. In general, there are four approaches to characterize the fatigue behavior of asphalt concrete. The methods including Phenomenological Approach, damage–energy fatigue approach, fatigue fracture mechanics and fatigue damage mechanics. Based on continuum damage mechanics, fatigue cycles of structural members in test can be greatly reduced and a unified approach of predicting fatigue crack initiation and fatigue crack propagation can be constructed. Generally, one damage variable is defined to account for the relative decrease of Young’s modulus with loading and, therefore, the change of Poisson ratio can not be considered. To accurately describe the coupling relationship between damage and stress fields, a bi-variable damage model to describe the variations of shear modulus and bulk modulus is the direction of future research.
APA, Harvard, Vancouver, ISO, and other styles
49

Jin, Ling Ling, Cai Yan Deng, Dong Po Wang, and Rui Ying Tian. "Research on Ultra-High Cycle Fatigue Property of 45 Steel." Advanced Materials Research 295-297 (July 2011): 1911–14. http://dx.doi.org/10.4028/www.scientific.net/amr.295-297.1911.

Full text
Abstract:
Fatigue property of 45 steel was studied in this paper with the method of ultrasonic fatigue testing, and SEM was used to analyze microscopic characteristics of the fatigue fracture. Fatigue test results show that: S-N curves descend continuously after 108 cycles, there is no fatigue limit as the traditional fatigue conception describes. Therefore, it is very dangerous to design welded structure working in the ultra-high cycle interval with the fatigue strength corresponding to 5×106 cycles. In the super-long life range, the fatigue property of welded joints is worse than the base metal. SEM analysis shows that: fatigue crack mainly initiates from the defects in the surface or sub-surface.
APA, Harvard, Vancouver, ISO, and other styles
50

Huang, Tsai-Wei, Chi-Huang Shih, Pai-Chien Chou, and Ting-Ling Chou. "Comparison between subjective and objective fatigue in patients with lung cancer and cancer-free participants: Evaluation of diagnostic criteria for cancer-related fatigue." Journal of Clinical Oncology 40, no. 28_suppl (October 1, 2022): 440. http://dx.doi.org/10.1200/jco.2022.40.28_suppl.440.

Full text
Abstract:
440 Background: Cancer-related fatigue is one of the most distressing symptoms of cancer patients. Finding a patient's fatigue often requires the use of a subjective assessment scale, such as the Brief Fatigue Inventory (BFI). There are few objective ways to measure fatigue. The aim of this study was to evaluate whether objective cancer-related fatigue (CRF) diagnostic criteria can differentiate cancer from cancer-free participants and to explore the relationship between subjective and objective fatigue. Methods: In this study, we used a photoplethysmography (PPG) smartband device to collect the PPG information from the patients. During PPG measurements, the heart rate variability (HRV) signal was included in the data collection process for subsequent calculation of the LF/HF ratio. Participants completed a self-report measure to assess demographics, fatigue levels, and brief sleep diaries for the next 7 days. At the same time, the participants wore a wristband to collect HRV signals, which triggered HRV every hour over a 24-hour period for 7 consecutive days. Results: Cancer patients (n = 71) and cancer-free controls (n = 75) were studied. Study participants completed CRF questionnaires and heart rate variability (HRV) records. The smartband can be worn continuously for 120 hours as an objective measure of activity phase, sleep phase and HRV. Compared with controls, cancer patients were more fatigued and were more likely to be disturbed by fatigue (t = -3.73, p < 0.001), especially general activity (t = -2.93, p < 0.001), walking ability (t = -3.6, p < 0.001) and normal work (t = -2.18, p = 0.03). Wearing the bracelet can get the LF/HF value of HRV in the time domain. The LF to HF and LF/HF disorder ratios were further divided into active and sleep phases. During sleep phase, the LF/HF disorder ratio was higher in the cancer group than in the control group regardless of whether fatigue was mild or moderate (t = -2.5, p = 0.01; t = -2.8, p = 0.01, respectively). The LF/HF disorder ratios was higher in cancer patients with mild fatigue than in cancer-free participants during sleep phase (40% vs 20%, respectively). Cancer patients with moderate fatigue had higher incidence of LF/HF disorder ratios than cancer-free participants (50% vs 20%, respectively). Conclusions: It can be concluded that the diagnostic criteria of subjective and objective CRF can distinguish cancer and cancer-free cases. Using this objective fatigue device could provide an indicator for further clinical monitoring. Clinical trial information: NCT04300842.
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography