Here’s how you know

  • U.S. Department of Health and Human Services
  • National Institutes of Health

NCCIH Clinical Digest

for health professionals

Mind and Body Approaches for Stress and Anxiety: What the Science Says

Clinical Guidelines, Scientific Literature, Info for Patients:  Mind and Body Approaches for Stress and Anxiety

yoga at home

.header_greentext{color:green!important;font-size:24px!important;font-weight:500!important;}.header_bluetext{color:blue!important;font-size:18px!important;font-weight:500!important;}.header_redtext{color:red!important;font-size:28px!important;font-weight:500!important;}.header_darkred{color:#803d2f!important;font-size:28px!important;font-weight:500!important;}.header_purpletext{color:purple!important;font-size:31px!important;font-weight:500!important;}.header_yellowtext{color:yellow!important;font-size:20px!important;font-weight:500!important;}.header_blacktext{color:black!important;font-size:22px!important;font-weight:500!important;}.header_whitetext{color:white!important;font-size:22px!important;font-weight:500!important;}.header_darkred{color:#803d2f!important;}.Green_Header{color:green!important;font-size:24px!important;font-weight:500!important;}.Blue_Header{color:blue!important;font-size:18px!important;font-weight:500!important;}.Red_Header{color:red!important;font-size:28px!important;font-weight:500!important;}.Purple_Header{color:purple!important;font-size:31px!important;font-weight:500!important;}.Yellow_Header{color:yellow!important;font-size:20px!important;font-weight:500!important;}.Black_Header{color:black!important;font-size:22px!important;font-weight:500!important;}.White_Header{color:white!important;font-size:22px!important;font-weight:500!important;} Relaxation Techniques

Relaxation techniques may be helpful in managing a variety of stress-related health conditions, including anxiety associated with ongoing health problems and in those who are having medical procedures. Evidence suggests that relaxation techniques may also provide some benefit for symptoms of post-traumatic stress disorder (PTSD) and may help reduce occupational stress in health care workers. For some of these conditions, relaxation techniques are used as an adjunct to other forms of treatment.

What Does the Research Show?

  • Biofeedback for anxiety and depression in children. A 2018 systematic review included 9 studies—278 participants total—on biofeedback for anxiety and depression in children and adolescents with long-term physical conditions such as chronic pain, asthma, cancer, and headache. The review found that, although biofeedback appears promising, at this point it can’t be recommended for clinical use in place of or in addition to current treatments. 
  • Heart rate variability biofeedback. A 2017 meta-analysis looked at 24 studies—484 participants total—on heart rate variability (HRV) biofeedback and general stress and anxiety. The meta-analysis found that HRV biofeedback is helpful for reducing self-reported stress and anxiety, and the researchers saw it as a promising approach with further development of wearable devices such as a fitness tracker.
  • Progressive muscle relaxation. A 2015 systematic review , which included two studies on progressive muscle relaxation in adults older than 60 years of age, with a total of 275 participants, found that progressive muscle relaxation was promising for reducing anxiety and depression. The positive effects for depression were maintained 14 weeks after treatment.
  • PTSD. A 2018 meta-analysis of 50 studies involving 2,801 participants found that relaxation therapy seemed to be less effective than cognitive behavioral therapy for PTSD and obsessive-compulsive disorder. No difference was found between relaxation therapy and cognitive behavioral therapy for other anxiety disorders, including generalized anxiety disorder, panic disorder, social anxiety disorder, and specific phobias. The review noted, however, that most studies had a high risk of bias, and there was a small number of studies for some of the individual disorders.
  • Anxiety in people with cancer. In the 2023 joint guideline issued by the Society for Integrative Oncology and the American Society for Clinical Oncology on integrative oncology care of symptoms of anxiety and depression in adults with cancer, relaxation therapies may be offered to people with cancer to improve anxiety symptoms during active treatment (Type: Evidence based; Quality of evidence: Intermediate; benefits outweigh harms; Strength of recommendation: Moderate). 
  • Relaxation techniques are generally considered safe for healthy people. In most research studies, there have been no reported negative side effects. However, occasionally, people report negative experiences such as increased anxiety, intrusive thoughts, or fear of losing control. 
  • There have been rare reports that certain relaxation techniques might cause or worsen symptoms in people with epilepsy or certain psychiatric conditions, or with a history of abuse or trauma. 

.header_greentext{color:green!important;font-size:24px!important;font-weight:500!important;}.header_bluetext{color:blue!important;font-size:18px!important;font-weight:500!important;}.header_redtext{color:red!important;font-size:28px!important;font-weight:500!important;}.header_darkred{color:#803d2f!important;font-size:28px!important;font-weight:500!important;}.header_purpletext{color:purple!important;font-size:31px!important;font-weight:500!important;}.header_yellowtext{color:yellow!important;font-size:20px!important;font-weight:500!important;}.header_blacktext{color:black!important;font-size:22px!important;font-weight:500!important;}.header_whitetext{color:white!important;font-size:22px!important;font-weight:500!important;}.header_darkred{color:#803d2f!important;}.Green_Header{color:green!important;font-size:24px!important;font-weight:500!important;}.Blue_Header{color:blue!important;font-size:18px!important;font-weight:500!important;}.Red_Header{color:red!important;font-size:28px!important;font-weight:500!important;}.Purple_Header{color:purple!important;font-size:31px!important;font-weight:500!important;}.Yellow_Header{color:yellow!important;font-size:20px!important;font-weight:500!important;}.Black_Header{color:black!important;font-size:22px!important;font-weight:500!important;}.White_Header{color:white!important;font-size:22px!important;font-weight:500!important;} Yoga, Tai Chi, and Qigong

A range of research has examined the relationship between exercise and depression. Results from a much smaller body of research suggest that exercise may also affect stress and anxiety symptoms. Even less certain is the role of yoga, tai chi, and qigong—for these and other psychological factors. But there is some limited evidence that yoga, as an adjunctive therapy, may be helpful for people with anxiety symptoms.

  • Yoga for children and adolescents. Findings from a 2021 meta-analysis and systematic review of 10 trials involving a total of 1,244 adolescents suggest a potential beneficial effect of tai chi and qigong on reducing anxiety and depression symptoms, and reducing cortisol level in adolescents. However, nonsignificant effects were found for stress, mood, and self-esteem. A  2020 systematic review  of 27 studies involving the effects of yoga on children and adolescents with varying health statuses, and with varying intervention characteristics, found that in studies assessing anxiety and depression, 58 percent showed reductions in both symptoms, while 25 percent showed reductions in anxiety only. Additionally, 70 percent of studies included in the review that assessed anxiety alone showed improvements. However, the reviewers noted that the studies included in the review were of weak-to-moderate methodological quality. 
  • Yoga, tai chi, and qigong for anxiety. A  2019 review  concluded that yoga as an adjunctive therapy facilitates treatment of anxiety disorders, particularly panic disorder. The review also found that tai chi and qigong may be helpful as adjunctive therapies for depression, but effects are inconsistent.
  • Yoga for anxiety. A  2021 randomized controlled trial examined whether Kundalini yoga and cognitive behavioral therapy (CBT) for generalized anxiety disorder (GAD) were each more effective than a control condition (stress education) and whether yoga was inferior to CBT for the treatment GAD. The trial found that Kundalini yoga was more efficacious for generalized anxiety disorder than the control, but the results support CBT remaining first-line treatment. A  2018 systematic review and meta-analysis  of 8 studies of yoga for anxiety (involving 319 participants with anxiety disorders or elevated levels of anxiety) found evidence that yoga might have short-term benefits in reducing the intensity of anxiety. However, when only people with diagnosed anxiety disorders were included in the analysis, there was no benefit. 
  • Yoga is generally considered a safe form of physical activity for healthy people when performed properly, under the guidance of a qualified instructor. However, as with other forms of physical activity, injuries can occur. The most common injuries are sprains and strains. Serious injuries are rare. The risk of injury associated with yoga is lower than that for higher impact physical activities.
  • Older people may need to be particularly cautious when practicing yoga. The rate of yoga-related injuries treated in emergency departments is higher in people age 65 and older than in younger adults.

.header_greentext{color:green!important;font-size:24px!important;font-weight:500!important;}.header_bluetext{color:blue!important;font-size:18px!important;font-weight:500!important;}.header_redtext{color:red!important;font-size:28px!important;font-weight:500!important;}.header_darkred{color:#803d2f!important;font-size:28px!important;font-weight:500!important;}.header_purpletext{color:purple!important;font-size:31px!important;font-weight:500!important;}.header_yellowtext{color:yellow!important;font-size:20px!important;font-weight:500!important;}.header_blacktext{color:black!important;font-size:22px!important;font-weight:500!important;}.header_whitetext{color:white!important;font-size:22px!important;font-weight:500!important;}.header_darkred{color:#803d2f!important;}.Green_Header{color:green!important;font-size:24px!important;font-weight:500!important;}.Blue_Header{color:blue!important;font-size:18px!important;font-weight:500!important;}.Red_Header{color:red!important;font-size:28px!important;font-weight:500!important;}.Purple_Header{color:purple!important;font-size:31px!important;font-weight:500!important;}.Yellow_Header{color:yellow!important;font-size:20px!important;font-weight:500!important;}.Black_Header{color:black!important;font-size:22px!important;font-weight:500!important;}.White_Header{color:white!important;font-size:22px!important;font-weight:500!important;} Meditation and Mindfulness-Based Stress Reduction

Some research suggests that practicing meditation may reduce blood pressure, anxiety and depression, and insomnia.

  • Mindfulness-based stress reduction. A  2023 randomized controlled trial involving 208 participants found that mindfulness-based stress reduction (MBSR) is noninferior to escitalopram, a commonly used first-line psychopharmacologic treatment for anxiety disorders. A  2021 randomized controlled trial of 108 adults with generalized social anxiety disorder found that cognitive behavioral group therapy and MBSR may be effective treatments with long-term benefits for patients with social anxiety networks that recruit cognitive and attention-regulation brain networks. The researchers noted that cognitive behavioral therapy and MBSR may both enhance reappraisal and acceptance emotion regulation strategies.
  • Mindfulness-based meditation. A  2019 review  concluded that as monotherapy or an adjunctive therapy, mindfulness-based meditation has positive effects on depression, and its effects can last for 6 months or more. Although positive findings are less common in people with anxiety disorders, the evidence supports adjunctive use. A 2019 analysis of 29 studies (3,274 total participants) showed that use of mindfulness-based practices among people with cancer significantly reduced psychological distress, fatigue, sleep disturbance, pain, and symptoms of anxiety and depression. However, most of the participants were women with breast cancer, so the effects may not be similar for other populations or other types of cancer. A  2014 meta-analysis  of 47 trials in 3,515 participants suggests that mindfulness meditation programs show moderate evidence of improving anxiety and depression. But the researchers found no evidence that meditation changed health-related behaviors affected by stress, such as substance abuse and sleep.
  • Mindfulness-based programs for workplace stress. A  2018 systematic review and meta-analysis  of nine studies examined mindfulness-based programs with an employee sample, which targeted workplace stress or work engagement, and measured a physiological outcome. The review found that mindfulness-based interventions may be a promising avenue for improving physiological indices of stress. 
  • Meditation is generally considered to be safe for healthy people.
  • A 2019 review found no apparent negative effects of mindfulness-based interventions and concluded that their general health benefits justify their use as adjunctive therapy for patients with anxiety disorders.

.header_greentext{color:green!important;font-size:24px!important;font-weight:500!important;}.header_bluetext{color:blue!important;font-size:18px!important;font-weight:500!important;}.header_redtext{color:red!important;font-size:28px!important;font-weight:500!important;}.header_darkred{color:#803d2f!important;font-size:28px!important;font-weight:500!important;}.header_purpletext{color:purple!important;font-size:31px!important;font-weight:500!important;}.header_yellowtext{color:yellow!important;font-size:20px!important;font-weight:500!important;}.header_blacktext{color:black!important;font-size:22px!important;font-weight:500!important;}.header_whitetext{color:white!important;font-size:22px!important;font-weight:500!important;}.header_darkred{color:#803d2f!important;}.Green_Header{color:green!important;font-size:24px!important;font-weight:500!important;}.Blue_Header{color:blue!important;font-size:18px!important;font-weight:500!important;}.Red_Header{color:red!important;font-size:28px!important;font-weight:500!important;}.Purple_Header{color:purple!important;font-size:31px!important;font-weight:500!important;}.Yellow_Header{color:yellow!important;font-size:20px!important;font-weight:500!important;}.Black_Header{color:black!important;font-size:22px!important;font-weight:500!important;}.White_Header{color:white!important;font-size:22px!important;font-weight:500!important;} Hypnotherapy

Hypnosis has been studied for anxiety related to medical or dental procedures. Some studies have had promising results, but the overall evidence is not conclusive.

  • A  2022 systematic review and meta-analysis of 19 trials found positive effects of hypnotherapy for reducing dental anxiety and fear during dental treatment. However, the reviewers noted that despite positive effects of hypnotic interventions in the systematic review, the results of the meta-analysis are very heterogeneous. 
  • The 2023 joint guideline issued by the Society for Integrative Oncology and the American Society for Clinical Oncology recommends that hypnosis may be offered to people with cancer to improve anxiety symptoms during cancer-related diagnostic and treatment procedures (Type: Evidence based; Quality of evidence: Intermediate; benefits outweigh harms; Strength of recommendation: Moderate).
  • Hypnosis is a safe technique when practiced by a trained, experienced, licensed health care provider.

.header_greentext{color:green!important;font-size:24px!important;font-weight:500!important;}.header_bluetext{color:blue!important;font-size:18px!important;font-weight:500!important;}.header_redtext{color:red!important;font-size:28px!important;font-weight:500!important;}.header_darkred{color:#803d2f!important;font-size:28px!important;font-weight:500!important;}.header_purpletext{color:purple!important;font-size:31px!important;font-weight:500!important;}.header_yellowtext{color:yellow!important;font-size:20px!important;font-weight:500!important;}.header_blacktext{color:black!important;font-size:22px!important;font-weight:500!important;}.header_whitetext{color:white!important;font-size:22px!important;font-weight:500!important;}.header_darkred{color:#803d2f!important;}.Green_Header{color:green!important;font-size:24px!important;font-weight:500!important;}.Blue_Header{color:blue!important;font-size:18px!important;font-weight:500!important;}.Red_Header{color:red!important;font-size:28px!important;font-weight:500!important;}.Purple_Header{color:purple!important;font-size:31px!important;font-weight:500!important;}.Yellow_Header{color:yellow!important;font-size:20px!important;font-weight:500!important;}.Black_Header{color:black!important;font-size:22px!important;font-weight:500!important;}.White_Header{color:white!important;font-size:22px!important;font-weight:500!important;} References

  • Carlson LE, Ismaila N, Addington EL, et al.  Integrative oncology care of symptoms of anxiety and depression in adults with cancer: Society for Integrative Oncology-ASCO guideline .  Journal of Clinical Oncology.  2023;41(28):4562-4591. 
  • Chugh-Gupta N, Baldassarre FG, Vrkljan BH.  A systematic review of yoga for state anxiety: considerations for occupational therapy . C anadian Journal of Occupational Therapy . 2013;80(3):150-170.
  • Cillessen L, Johannsen M, Speckens AEM, et al . Mindfulness-based interventions for psychological and physical health outcomes in cancer patients and survivors: a systematic review and meta-analysis of randomized controlled trials .  Psychooncology . 2019;28(12):2257-2269. 
  • Cramer H, Lauche R, Anheyer D, et al.  Yoga for anxiety: a systematic review and meta-analysis of randomized controlled trials .  Depress Anxiety . 2018;35(9):830-843.
  • Goessl VC, Curtiss JE, Hofmann SG.  The effect of heart rate variability of biofeedback training on stress and anxiety: a meta-analysis .  Psychological Medicine . 2017;47(15):2578-2586.
  • Goldin PR, Thurston M, Allende S, et al . Evaluation of cognitive behavioral therapy vs mindfulness meditation in brain changes during reappraisal and acceptance among patients with social anxiety disorder: a randomized clinical trial .  JAMA Psychiatry . 2021;78(10):1134-1142.
  • Goyal M, Singh S, Sibinga EMS, et al.  Meditation programs for psychological stress and well-being: a systematic review and meta-analysis.   JAMA Internal Medicine . 2014;174(3):357-368.
  • Greenlee H, Balneaves LG, Carlson LE, et al.  Clinical practice guidelines on the use of integrative therapies as supportive care in patients treated for breast cancer .  Journal of the National Cancer Institute Monographs.  2014;50:346-358.
  • Heckenberg RA, Eddy P, Kent S, et al.  Do workplace-based mindfulness meditation programs improve physiological indices of stress? A systematic review and meta-analysis .  Journal of Psychosomatic Research.  2018;114:62-71.
  • Hoge EA, Bui E, Mete M, et al.  Mindfulness-based stress reduction vs escitalopram for the treatment of adults with anxiety disorders: a randomized clinical trial .  JAMA Psychiatry . 2023;80(1):13-21.
  • James-Palmer A, Anderson EZ, Zucker L, et al. Yoga as an intervention for the reduction of symptoms of anxiety and depression in children and adolescents: a systematic review .  Frontiers in Pediatrics . 2020;8:78.
  • Liu X, Li R, Cui J, et al.  The effects of tai chi and qigong exercise on psychological status in adolescents: a systematic review and meta-analysis .  Frontiers in Psychology . 2021;12:746975.
  • Klainin-Yobas P, Oo WN, Suzanne Yew PY, et al.  Effects of relaxation interventions on depression and anxiety among older adults: a systematic review .  Aging and Mental Health . 2015;19(12):1043-1055.
  • Montero-Marin J, Garcia-Campayo J, López-Montoyo A, et al.  Is cognitive-behavioural therapy more effective than relaxation therapy in the treatment of anxiety disorders? A meta-analysis .  Psychological Medicine . 2018;48(9):1427-1436.
  • Saeed SA, Cunningham K, Bloch RM.  Depression and anxiety disorders: benefits of exercise, yoga, and meditation .  American Family Physician . 2019;99(10):620-627.
  • Simon NM, Hofmann SG, Rosenfield D, et al.  Efficacy of yoga vs cognitive behavioral therapy vs stress education for the treatment of generalized anxiety disorder: a randomized clinical trial .  JAMA Psychiatry . 2021;78(1):13-20.
  • Thabrew H, Ruppeldt P, Sollers JJ 3rd.  Systematic review of biofeedback interventions for addressing anxiety and depression in children and adolescents with long-term physical conditions .  Applied Psychophysiology and Biofeedback . 2018;43(3):179-192.
  • Wolf TG, Schläppi S, Benz CI, et al.  Efficacy of hypnosis on dental anxiety and phobia: a systematic review and meta-analysis .  Brain Sciences . 2022;12(5):521.

NCCIH Clinical Digest is a service of the National Center for Complementary and Integrative Health, NIH, DHHS. NCCIH Clinical Digest, a monthly e-newsletter, offers evidence-based information on complementary health approaches, including scientific literature searches, summaries of NCCIH-funded research, fact sheets for patients, and more.

The National Center for Complementary and Integrative Health is dedicated to exploring complementary health products and practices in the context of rigorous science, training complementary health researchers, and disseminating authoritative information to the public and professionals. For additional information, call NCCIH’s Clearinghouse toll-free at 1-888-644-6226, or visit the NCCIH website at nccih.nih.gov . NCCIH is 1 of 27 institutes and centers at the National Institutes of Health, the Federal focal point for medical research in the United States.

Content is in the public domain and may be reprinted, except if marked as copyrighted (©). Please credit the National Center for Complementary and Integrative Health as the source. All copyrighted material is the property of its respective owners and may not be reprinted without their permission.

Subscriptions

NCCIH Clinical Digest is a monthly e-newsletter that offers evidence-based information on complementary and integrative health practices.

Clinical Digest Archive

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Open access
  • Published: 27 November 2021

Psychological and biological resilience modulates the effects of stress on epigenetic aging

  • Zachary M. Harvanek   ORCID: orcid.org/0000-0003-3702-1051 1 ,
  • Nia Fogelman 2 ,
  • Ke Xu   ORCID: orcid.org/0000-0002-6472-7052 1 , 3 &
  • Rajita Sinha   ORCID: orcid.org/0000-0003-3012-4349 1 , 2 , 4 , 5  

Translational Psychiatry volume  11 , Article number:  601 ( 2021 ) Cite this article

23k Accesses

46 Citations

417 Altmetric

Metrics details

Our society is experiencing more stress than ever before, leading to both negative psychiatric and physical outcomes. Chronic stress is linked to negative long-term health consequences, raising the possibility that stress is related to accelerated aging. In this study, we examine whether resilience factors affect stress-associated biological age acceleration. Recently developed “epigenetic clocks” such as GrimAge have shown utility in predicting biological age and mortality. Here, we assessed the impact of cumulative stress, stress physiology, and resilience on accelerated aging in a community sample ( N  = 444). Cumulative stress was associated with accelerated GrimAge ( P  = 0.0388) and stress-related physiologic measures of adrenal sensitivity (Cortisol/ACTH ratio) and insulin resistance (HOMA). After controlling for demographic and behavioral factors, HOMA correlated with accelerated GrimAge ( P  = 0.0186). Remarkably, psychological resilience factors of emotion regulation and self-control moderated these relationships. Emotion regulation moderated the association between stress and aging ( P  = 8.82e−4) such that with worse emotion regulation, there was greater stress-related age acceleration, while stronger emotion regulation prevented any significant effect of stress on GrimAge. Self-control moderated the relationship between stress and insulin resistance ( P  = 0.00732), with high self-control blunting this relationship. In the final model, in those with poor emotion regulation, cumulative stress continued to predict additional GrimAge Acceleration even while accounting for demographic, physiologic, and behavioral covariates. These results demonstrate that cumulative stress is associated with epigenetic aging in a healthy population, and these associations are modified by biobehavioral resilience factors.

Similar content being viewed by others

research studies on stress

Associations of stress and stress-related psychiatric disorders with GrimAge acceleration: review and suggestions for future work

research studies on stress

Epigenetic aging and perceived psychological stress in old age

research studies on stress

Trauma, adversity, and biological aging: behavioral mechanisms relevant to treatment and theory

Introduction.

Cumulative stress can have adverse psychiatric and physical effects, increasing risk for cardiometabolic diseases, mood disorders, post-traumatic stress disorder and addiction [ 1 , 2 , 3 , 4 , 5 , 6 , 7 , 8 , 9 , 10 , 11 ]. There are several potential psychological and biological mechanisms through which these effects may occur. For example, stress may reduce psychological resilience measures such as emotion regulation and self-control that are known to protect against psychiatric and physical health outcomes [ 1 , 12 , 13 , 14 ]. Notably, emotional stress exposure decreases cognitive and emotion regulation abilities [ 15 , 16 , 17 , 18 ], and this effect may be modulated by cortisol [ 15 ]. Furthermore, stress decreases self-control abilities [ 19 , 20 , 21 ] and impacts the likelihood of individuals engaging in healthy behaviors such as exercise or maintaining a healthy diet, and is associated with unhealthy behaviors such as smoking, alcohol, and drug use [ 22 , 23 , 24 , 25 ]. Recent evidence also suggests that stress effects on metabolic health may be affected by BMI-related changes in insulin resistance and other gut hormones [ 26 , 27 ]. Indeed, stress’s effects on physiology resulting in alterations in neuro-hormonal signaling pathways as well as increased inflammation are well documented [ 26 , 28 , 29 , 30 ]. Both stress and these physiologic changes may increase the risk of multiple physical and psychiatric illnesses, which in turn increase morbidity and mortality risk. This has often been described as an increased allostatic load, and notably many of these processes, such as metabolic and cardiovascular dysfunction, have been associated with human aging [ 31 ]. For example, insulin signaling might be linked to aging and aging-related diseases in humans [ 32 ], with recent data on metformin (a treatment for insulin resistance) suggesting it may be useful as an anti-aging drug [ 33 ].

There is growing evidence that cumulative stress may adversely impact health via accelerating the cellular aging process. For example, stress shortens telomere length and alters telomerase activity, and this interaction is modified by behavioral and psychological resilience factors [ 34 , 35 , 36 , 37 ]. However, recent studies have demonstrated mixed results on whether characteristics that contribute to resilience improve or worsen the impact of stress on health [ 38 , 39 , 40 , 41 , 42 , 43 , 44 , 45 , 46 , 47 ]. These data suggest that resiliency factors may modulate the relationship between chronic stress and aging, but to our knowledge this has not been tested in a healthy community sample. While there are many aspects of resilience, including cultural/societal, environmental, and personal which can decrease the negative consequences of stressors on individuals, herein we will focus on personal-level, psychological skills, including self-control and emotion regulation.

Recently developed DNA methylation-based epigenetic “clocks” appear to provide a more accurate measure of biological age than telomere length [ 48 , 49 , 50 , 51 ]. These clocks are built from a set of DNA methylation markers that correlate with chronologic age and serve as molecular estimators of biological age in cells, tissues, and individuals [ 52 ]. Epigenetic clocks have a significantly higher predictive value than previously used measures such as telomere length for frailty, [ 53 ] mortality risk [ 54 , 55 ], hazard ratios [ 56 ], and chronologic age [ 57 ]. The development of these biological aging markers promises to not only aid in identifying individuals at higher risk for aging-related illnesses, but potentially also developing interventions to prevent accelerated aging.

Previous studies (reviewed by Palma-Gudiel et al [ 58 ]) have utilized epigenetic clocks to demonstrate associations between trauma, early life adversity, or low socioeconomic status and accelerated epigenetic aging. Studies have often been focused upon selected populations, such military veterans [ 45 ], individuals with significant trauma histories [ 59 ], or specific cohorts at higher risk [ 60 , 61 , 62 ]. Notably, these studies did not exclude, and often explicitly included, individuals with significant mental and physical illnesses, including PTSD, MDD, and other disabilities [ 59 , 63 ]. These studies also primarily utilized epigenetic clocks trained upon chronologic age. However, a recently developed epigenetic clock, GrimAge, was trained using biomarkers of mortality and indicators of health, and has superior performance in predicting health outcomes when compared with other epigenetic clocks [ 51 , 64 ].

We utilized GrimAge Acceleration (“GAA”, defined as the residual of the regression of GrimAge to chronologic age, with a positive number indicating biological age greater than chronologic age) to conduct a cross-sectional study to answer three questions. First, is cumulative stress related to epigenetic markers of biological aging in a healthy young-to-middle-aged community population? Second, if stress is associated with epigenetic aging, does stress-related physiology contribute to stress-associated biological aging? And finally, how do psychological factors that contribute to resilience modulate these relationships? Based on previous research, we hypothesized that cumulative stress will be positively associated with GrimAge Acceleration (GAA), that stress effects on GrimAge will be related to changes in the hypothalamic-pituitary-adrenal axis (HPA) and insulin sensitivity, and that strong emotion regulation as measured by the Difficulties in Emotion Regulation Scale (DERS, [ 65 ]) and high self-control as measured by the Self Control Scale-Brief (SCS-B, [ 66 ]) will moderate the relationships between stress, physiology, and accelerated aging (See Fig. 1 for a model summarizing our hypotheses).

figure 1

We hypothesize that stress is positively associated with accelerated biological aging, which we measure via GrimAge Acceleration (GAA), and that this relationship will be mediated by stress-related physiologic changes such as insulin and HPA signaling. We also hypothesize that strong psychological resilience factors will be protective against the negative consequences of stress on aging. Note that these relationships are predictive, not causative, as this study is cross-sectional and thus directionality of relationships cannot be conclusively examined.

Materials and methods

Cohort recruitment.

The participant cohort included 444 community adults between the ages of 18–50 in the greater New Haven, CT area who volunteered to participate in a study examining the role of stress and self-control at the Yale Stress Center as previously described [ 67 ]. Briefly, participants were recruited via advertisements online, in local newspapers, and at a community center between 2008 and 2012. Participants were excluded if they had a substance use disorder (not including nicotine) as assessed via the Structured Clinical Interview for Diagnostic and Statistical Manual of Mental Disorders, 4th Edition (SCID-I for DSM-IVTR), were pregnant, had a chronic medical condition (e.g, hypertension, diabetes, hypothyroidism), or were unable to read English at or above the 6th grade level. Participants were also excluded if they had a concussion with loss of consciousness greater than 30 minutes, another head injury such as documented traumatic brain injury or another injury with documented lasting deficits, or were using any prescribed medications for any psychiatric or medical disorders. Breathalyzer and urine toxicology screens were conducted at each appointment to ensure the participants were drug-free. Of a total of 1000 potential participants who underwent initial screening for eligibility, epigenetic data combined with physiologic and behavioral data were available on 444, who comprised the current sample. All participants provided written and verbal informed consent to participate, and the research protocol was reviewed and approved by the Yale IRB.

Initial assessment and measurement of physiologic parameters

All eligible subjects met with a research assistant for two intake sessions to complete a physical health review with the Cornell Medical Index (CMI, [ 68 ]), structured clinical interview for diagnoses (SCID) of DSM-IVTR psychiatric illnesses, cumulative stress interview, self-report assessments and a separate morning biochemical evaluation after fasting overnight. The structured clinical interview was performed by masters’ or doctoral level clinical research staff. Fasting insulin and glucose were obtained and Cortisol was assessed at four time-points, spaced 15 min apart beginning at 7:30 AM after overnight fasting and collected while participants were in a quiet and comfortable laboratory setting at the Yale Stress Center. Participants were financially compensated for participating in the study.

Psychological measures

Cumulative stress was assessed using the Cumulative Adversity Inventory (CAI, [ 69 ]), a 140-item multifaceted interview-based assessment of life events and subjective stress through which trained interviewers asked participants about specific stressful events that occurred during their lifetime, which comprised the subscales of major life events, life trauma events and recent life events. For purposes of scoring, a “yes” to the specific stressful event occurring led to a “1” and a sum of all the “yes” endorsements comprised the subscale score for these events subscale. The final subscale of chronic stress was the participant’s own sense of feeling overwhelmed and unable to manage the events for the other subscales listed. This was rated on a “not true”, “somewhat true”, or “very true” scale, with assigned scores of 0, 1, and 2, respectively. The final score is a sum of these values for the chronic stress subscale. The CAI-total score was a sum of each of the subscale score with a higher score indicating a higher overall level of lifetime cumulative stress. The CAI has been demonstrated to have excellent overall reliability as reported in previous research [ 12 , 26 , 70 , 71 , 72 ]. In our population for this study, the alpha reliability is 0.86. It has been previously shown to predict cumulative stress related brain volume reductions and sensitized stress functional responses as well as prediction of physical, metabolic and behavioral responses [ 26 , 70 , 71 , 72 ].

Emotion regulation was assessed using the Difficulties with Emotion Regulation Scale (DERS, [ 65 ]), which is a 41-item trait-level measure that assesses across domains of lack of emotional awareness, goals, clarity, strategies, acceptance, and impulse control in managing emotions. Higher scores on the DERS correspond to lower ability to regulate emotion. Alpha reliability has been reported to be >0.90 for the total score, and ≥0.80 for the sub-scores [ 65 ]. In this population, the alpha reliability is 0.92.

Self-control was assessed using the Self-Control Survey-Brief (SCS-B, [ 66 ]), which is a 13-item scale that assesses overall self-control. A higher score on the SCS-B suggests a stronger level of self-control. There are no sub-scores provided by the SCS-B, and the overall SCS-B has been reported to have an alpha reliability >0.80 [ 66 ]. The alpha reliability in this study is 0.85.

The Cornell Medical Index (CMI) was used to assess for participants’ current health. It is a 195-question interview that captures both physical and psychological health symptoms, and has been validated as an indicator for current general health in many studies [ 68 , 73 , 74 ]. A higher score on the CMI suggests more symptoms and worse overall health. The alpha reliability of the total CMI is 0.94. The psychological subscore has an alpha reliability of 0.92, and the biological subscore has a reliability of 0.90.

Cronbach alpha reliabilities for each of the scales described above were obtained using the alpha function in the R psych package [ 75 ].

DNA methylation and epigenetic clock analysis

DNA for epigenetic analysis was collected from whole blood samples as previously described [ 67 ]. Briefly, all samples were profiled using Illumina Infinium HumanMethylation450 Beadchips, which covers 96% of CpG islands and 99% of RefSeq genes. Quality control on these data are as previously published [ 67 ]. They are described in brief below:

Probe QC : To ensure high-quality data, we set a more stringent threshold of P  < 10 –12 . Intensity values showing P  > 10 −12 were set as zero. Additionally, we removed 11,648 probes on sex chromosomes and 36,535 probes within 10 base pairs of single-nucleotide polymorphisms. Finally, a total of 47,791 probes were removed and the remaining 437,722 probes were used for further analysis.

Sample QC : Using a detection P value < 10 –12 , one sample showing a call rate < 98% was excluded from analysis. Five samples showing sex discrepancy between the methylation predicted sex and self-reported sex were also excluded from analysis.

Data processing and normalization : Data processing and normalization were performed using the recently published protocol (Lehne et al., 2015). We first perform background correction and within-array normalization to the original green/red channel intensity data using the preprocessIllumina function in the minfi R package. The processed data were transformed to M/U methylation categories. Next, we separately performed between-array-normalization with the quantile method using the normalizeBetweenArrays function in the limma R package (version 3.26.2) after dividing the data matrix into 6 independent parts: Type I M Green, Type I M Red, Type I U Red, Type I U Green, Type II Red, Type II Green. The normalized data were merged and the beta value at each CpG site was determined.

After obtaining beta values, epigenetic clock analysis was performed as described in Lu et al. using the New Methylation Age Calculator at https://dnamage.genetics.ucla.edu/new [ 51 ]. Data were normalized as per their protocol, and the advanced analysis option was used. We focus on GrimAge acceleration (GAA), which is defined as the residuals of a linear correlation of GrimAge to chronologic age. No effects of array batch on GAA were observed (Supplementary Fig. 1 ).

The analyses herein were performed without accounting for individual variations in cell types. The Houseman method was used to determine cell type proportion [ 76 ], and the inclusion of cell fractions as covariates in a linear model does not impact the primary conclusions of this paper (see Supplementary material).

Statistical analysis

Data organization and analysis were conducted using R 3.6.3 [ 77 ] and RStudio. Linear regressions were first implemented to examine univariate associations between independent and dependent variables. Multivariable linear regressions adjust for demographic (sex, race, years of education, marital status, income) and behavioral (smoking, alcohol use, and BMI) covariates unless otherwise stated. These covariates were selected due to prior work demonstrating a relationship to epigenetic aging. Chronologic age is incorporated into the model as part of the calculation of GAA (the residual of GrimAge regressed upon chronologic age). There was no significant correlation between chronologic age and GAA. Analyses of the relationship between CAI, GAA, psychological and physiologic variables were performed in both the univariate unadjusted model and the multivariate adjusted model accounting for demographic and behavioral measures, but except when the conclusions differ, statistical values in the text represent the multivariate models for simplicity. CAI, DERS, and SCS were mean-centered to address issues of collinearity (particularly regarding individual regression coefficients) when assessing for moderation.

All tests were two-tailed with alpha set at 0.05. Statistical significance in both standard linear regressions and moderation analyses were assessed from t values. R 2 reported on plots represent the simple relationship between the stated variables, while adjusted R 2 values in the text represent the model. Partial η 2 values represent the effect size for the specific variable, with a value >= 0.01 typically indicating a small effect, >= 0.06 a medium effect, and >= 0.14 a large effect [ 78 ]. Wilcoxon signed-rank test was used to compare data between sexes. Mediation analysis was performed to determine if stress impacts GAA via behavioral and physiologic factors. Simple mediation effects were calculated via R using 10,000 simulations without bootstrapping using the mediation package [ 79 ]. Mediation was considered significant if the proportion mediated was greater than 0 with an alpha of 0.05. Serial mediation was calculated via R using the Lavaan package [ 71 ], with an indirect effect defined as the product of the coefficients of the effect of stress on BMI, of BMI on HOMA, and of HOMA on GAA. Assessment of the individual variables’ attributable GrimAge acceleration as well as confidence intervals were calculated using the Emmeans package using unadjusted pairwise comparisons.

Demographics and clinical characteristics

As shown in Table 1 , study participants were healthy and without evidence of medical or psychiatric diseases. The majority were non-smokers (79.6%), social drinkers with low risky alcohol intake screening scores (72.7% of participants have Alcohol Use Disorders Identification Test (AUDIT) < 8, and 91.7% < 15), and were not obese (74.5% of participants have a BMI < 30, 89.2% < 35). Both physical and psychological symptoms assessed on the Cornell Medical Index (CMI, [ 68 ]) were low, with 86% of participants scoring below the typical screening threshold of 30.

Cumulative stress predicts accelerated biological aging as measured by GrimAge

As expected, there was a high association between individuals’ chronologic age and GrimAge (Age: t  = 51.4, P  < 2e−16, adjusted R 2  = 0.856, Fig. 2A ). This relationship is not altered by inclusion of the covariates of smoking, alcohol use, BMI, race, sex, income, and years of education (Age: t  = 49.1, P  < 2e−16, partial η 2  = 0.848; model (GrimAge ~ Age + covariates) adjusted R 2  = 0.912), and this relationship remained significant accounting for cellular fractions (Supplementary Table 1 ). Also, using a univariate linear regression, greater cumulative stress as measured by the total Cumulative Adversity Index (CAI) score significantly predicted higher GAA (CAI: t  = 4.82  P  = 2.00e−6, η 2  = 0.050, adjusted R 2  = 0.0478, Fig. 2B ). While there were significant differences in GAA based on sex ( P  = 1.33e−7), both males (CAI: P  = 3.35e−4, adjusted R 2  = 0.0586) and females (CAI: P  = 3.12e−5, adjusted R 2  = 0.0652) demonstrated similar correlations between stress and GAA. Further analysis showed these results are consistent across CAI subscales, as well as with the Childhood Trauma Questionnaire and several of its subscales (Supplementary Table 2 ).

figure 2

A Chronologic age significantly predicts GrimAge ( P  < 2e−16). B Cumulative stress total as measured by the CAI (CAI-Total) significantly predicts GAA before ( P  = 2.00e−6) and after accounting for covariates. C Higher insulin resistance (as measured by HOMA) shows a significant positive correlation with GAA before ( P  = 1.11e−8) and after accounting for covariates. D The Cortisol/ACTH ratio is negatively correlated with GAA before accounting for covariates ( P  = 2.39e−6), but not afterward. P and R 2 values in the figure represent simple univariate models (Y ~ X). In the main text, models are adjusted for covariates as stated.

After accounting for the covariates of smoking, alcohol use, BMI, race, sex, income, and years of education, the relationship between GAA and CAI remains significant (CAI: t  = 2.073, P  = 0.0388, partial η 2  = 0.010; model (GAA ~ CAI-total + covariates): adjusted R 2  = 0.3869); individual covariate effects shown in Supplementary Table 3 ). When considered as potential mediators of the relationship between stress and GAA, BMI (proportion mediated = 0.288, P  = 0.0042) and smoking (proportion mediated = 0.443, P  = 0.0030), but not alcohol use (proportion mediated = 0.001, P  = 0.931), show partial mediating effects (Supplementary Table 4 ).

Consistent with the underlying assumption that GAA is related to measures of health, GAA also predicted psychological and physical health symptoms as measured by the CMI (Supplementary Fig. 2A ; total CMI: t  = 3.449, P  = 6.18e−4, adjusted R 2  = 0.024).

Stress-related physiology is associated with GrimAge acceleration

Given the known relationship between cumulative stress and physiology, we assessed the relationship between the stress-related physiologic factors of insulin resistance and HPA-axis signaling and GAA. We found that higher HOMA (a measure of insulin resistance) significantly predicted GAA (Fig. 2C , HOMA: t  = 2.362, P  = 0.0186, partial η 2  = 0.013; model (GAA ~ HOMA + Covariates): adjusted R 2  = 0.389).

We then assessed whether cortisol/ACTH ratio changes impacted GAA. Indeed, low cortisol/ACTH ratio, a measure of adrenal sensitivity, was associated with GAA in a simple univariate model, (Fig. 2D , Cort/ACTH ratio: t  = −4.78, P  = 2.39e−6, η 2  = 0.049, adjusted R 2  = 0.0470), though this becomes non-significant when accounting for covariates (Cort/ACTH ratio: t  = −0.721, P  = 0.471, partial η 2  = 0.001; model (GAA ~ Cort/ACTH + Covariates): adjusted R 2  = 0.3816). We also find a significant association between stress and Cortisol/ACTH ratio (Supplementary Fig. 2B , CAI: t  = −2.146  P  = 0.0324; model (Cort/ACTH ratio ~ CAI + covariates): adjusted R 2  = 0.2197).

Emotion regulation moderates the relationship between stress and GrimAge acceleration directly

We then asked whether the relationship between cumulative stress and epigenetic aging was modulated by characteristics that contribute to an individual’s psychological resilience. We hypothesized that strong emotion regulation abilities would be protective against stress-related accelerated aging. We found that emotion regulation as assessed by the Difficulties in Emotion Regulation Scale (DERS, [ 65 ]) significantly moderated the relationship between GAA and CAI (Fig. 3A , CAI:DERS: F  = 11.22, P  = 8.82e−4, partial η 2  = 0.025; model (GAA ~ CAI X DERS + covariates): adjusted R 2  = 0.4004), such that poor emotion regulation significantly increased the effects of CAI on GAA. There was not a significant difference between males and females in emotion regulation ( P  = 0.0949).

figure 3

A Individuals with stronger emotion regulation (as measured by lower DERS scores) suffer less GAA at high stress than individuals with poor emotion regulation before (GAA ~ CAI X DERS P  = 9.51e−5; GAA ~ CAI X DERS + Covariates: P  = 8.82e−4) and after accounting for covariates. For panel A, “good” represents the slope at the 25th percentile of DERS, “fair” at the 50th percentile, and “poor” the 75th percentile. B Better self-control (as measured by higher B-SCS scores) is protective against the effects of stress on GAA before accounting for covariates (GAA ~ CAI X SCS P  = 0.00226; GAA ~ CAI X SCS + Covariates: P  = 0.130), but not after including them in the model. C Stronger self-control moderates the relationship between stress and insulin resistance before (HOMA ~ CAI X SCS P  = 0.0115; HOMA ~ CAI X SCS + Covariates P  = 0.00732) and after accounting for covariates. For panels (B) and (C), “good” represents the slope at the 75th percentile of B-SCS, “fair” at the 50th percentile, and “poor” the 25th percentile.

Self-control moderates the association between stress and insulin resistance, which is associated with GrimAge acceleration

We next assessed whether psychological resilience in the form of self-control (as measured via the SCS-B, [ 66 ]) alters the association between cumulative stress and GAA. We found higher self-control is protective against the effects of stress on GAA before accounting for covariates, but the interaction became non-significant when covariates were accounted for (Fig. 3B , CAI:SCS: F  = 2.303, P  = 0.130, partial η 2  = 0.005; model (GAA ~ CAI X SCS + Covariates: adjusted R 2  = 0.3874).

Given the potential interplay between self-control, insulin resistance, and stress, we next asked whether self-control moderated the relationship between stress and HOMA. We observed that, even when covariates are accounted for, self-control moderates the positive relationship between stress and HOMA, with stronger self-control blunting their relationship (Fig. 3C , CAI:SCS: F = 7.263, P  = 0.00732, partial η 2  = 0.017; model (HOMA ~ CAI X SCS + Covariates: adjusted R 2  = 0.2871). Notably, self-control does not moderate the relationship between CAI and BMI (CAI:SCS: F  = 0.679, P  = 0.41). Self-control did not significantly differ between males and females ( P  = 0.0550).

Exploratory mediation analyses suggest stress influences GrimAge via BMI and HOMA

While our ability to draw causative inferences are limited by the cross-sectional nature of our data, we used mediation analyses to explore potential relationships between weight, insulin resistance, and GAA. We hypothesized that the effects of BMI on GAA may be mediated through insulin resistance. Indeed, mediation analysis suggested that a significant portion of the effect of BMI on GAA may be mediated through HOMA (Supplementary Fig. 3A , proportion mediated = 0.247, P  = 0.02). Given these findings, we next asked whether BMI and insulin resistance act sequentially to mediate the effects of stress on GAA. We identified a significant indirect effect, suggesting that stress may affect GAA through increased BMI and elevated insulin resistance (Supplementary Fig. 3B , indirect effect = 0.003; P  = 0.030), though there continues to be a significant direct effect of stress on GAA as well (direct effect = 0.034, P  = 0.009).

Cumulative stress and estimated change in GrimAge

Finally, we sought to identify the comparative contributions of our significant variables to GAA. To do this, we constructed a linear regression model using all demographic covariates (sex, race, marital status, education, income), behavioral covariates (smoking, alcohol, BMI), physiologic factors (HOMA, Cortisol/ACTH ratio), and psychological factors. In this model, we continue to see a significant interaction between stress and emotion regulation in relation to GAA (CAI:DERS t  = 3.424, P  = 0.000677, partial η 2  = 0.027; model (GAA ~ CAI-total X DERS + HOMA + Cort/ACTH ratio + SCS + Covariates): adjusted R 2  = 0.4056). Notably in this model, HOMA ( t  = 2.308, P  = 0.0215, partial η 2  = 0.012), BMI ( t  = 2.641, P  = 0.00857, partial η 2  = 0.016), and smoking ( t  = 10.47, P  < 2e−16, partial η 2  = 0.204) also demonstrate significant effects on GAA. The impact of the cortisol/ACTH ratio on GAA is not significant ( t  = −0.668, P  = 0.504, partial η 2  = 0.001), and its removal from the model does not impact any of the above conclusions.

Using this final linear model, we estimated the changes in GrimAge for each significant variable (Table 2 ) using estimated marginal means [ 80 ]. When comparing the effects of high stress (CAI-total: 75th percentile) versus low stress (CAI-total: 25th percentile) in those with poor emotion regulation (DERS: 75th percentile), stress was associated with half a year of aging independent of all other covariates and physiologic factors. However, when emotion regulation was strong (DERS: 25th percentile), stress did not independently predict GAA. Again comparing 75th versus 25th percentiles, BMI independently was related to an increase of 0.46 years of GrimAge, and HOMA for ¼ of a year. We also identified daily smoking (3.8 years), male sex (1.2 years), self-identifying as Black (1 year), and never having married (0.71 years) as covariates that significantly predicted accelerated GrimAge. When accounting for cellular fractions we see similar results regarding the relationships between stress, emotion regulation, and GAA. However, when accounting for cellular fractions, the associations between GAA and both HOMA and marital status become non-significant (Supplementary Table 5 ). Prior literature [ 51 ] suggests that GrimAge predicts the hazard ratio exponentially (HR = 1.1 GAA ). Thus, each additional year of GAA would be expected to increase the relative risk of death by approximately 10%.

In this study, we report novel findings that cumulative stress is associated with accelerated epigenetic aging in a healthy, young-to-middle-aged community sample, even after adjusting for sex, race, BMI, smoking, alcohol use, income, marital status, and education. Epigenetic aging was measured by GrimAge, a marker which has previously been associated with increased morbidity and mortality and correlates with physical and psychological health symptoms in our study. The relationship between stress and age acceleration is most prominent in those with poor emotion regulation and was related to behavioral factors such as smoking and BMI. Both stress and GAA were associated with changes in insulin resistance, which was moderated via self-control. These results suggest a relationship between stress, physiology, and accelerated aging that is moderated by emotion regulation and self-control. Overall, these findings point to multiple potentially modifiable biobehavioral targets of intervention that may reduce or prevent the deleterious effects of stress on aging and long-term health outcomes.

This study included a generally healthy, young-to-middle-aged community population, yet we still identified a significant relationship between cumulative stress and age acceleration. The population was taking no prescription medications for any medical conditions, nor were they suffering from current mental illnesses, including major depressive disorder or generalized anxiety disorder. The study includes individuals with obesity, as well as a small number of individuals with risky drinking levels as determined by the AUDIT scores. The frequency of these individuals in the sample is generally in line with those in a community population, and thus we included alcohol use and BMI as covariates to account for the impact of these variables on the results. Prior work has demonstrated that GrimAge better predicts mortality than other epigenetic clocks, and GrimAge predicts lifespan more accurately than self-reporting smoking history, demonstrating that GrimAge is a biologically meaningful and potentially clinically useful biomarker for health [ 51 , 64 ]. Our findings are consistent with recent work showing that those with significant trauma histories [ 59 , 81 ] or with diagnoses of mental illnesses, such as Bipolar disorder or MDD, may experience accelerated aging as measured by epigenetic clocks [ 57 , 81 , 82 , 83 , 84 ]. In particular, this study builds on previous findings by Zannas et al that demonstrated a relationship between trauma and epigenetic aging using the Horvath clock. However, to the best of our knowledge this is the first study to investigate the impact of cumulative stress on epigenetic aging in a healthy community sample without significant physical or mental illness. Also it is the first to our knowledge to identify factors that contribute to psychological resilience as potential modulators of such an effect. This opens the possibility that the distinction between the effects of stress on pathologic and non-pathologic samples may be along a continuum. It would be interesting to examine resilience characteristics in the population studied by Zannas et al to determine if there is a limit to the protective effects of psychological resilience. Thus, preventive interventions that decrease stress and improve resilience may be useful for maintaining long-term mental and physical health.

The relationship between stress and epigenetic aging appears to be modulated via specific psychological traits, including emotion regulation and self-control. Those with better emotion regulation and higher levels of self-control were observed to have less age acceleration even at similar levels of stress. Indeed, based on their GAA, our estimates indicate that the relationship between stress and GrimAge is as powerful as BMI, but only for those with poor emotion regulation. As these are skills that may be developed through specific psychological interventions [ 85 ], these results raise the possibility that building emotion regulation skills could result in improvements in epigenetic aging, morbidity, and mortality [ 86 ] for these populations. As this is a cross-sectional study, we are not able to address whether these relationships are causal. These novel cross-sectional findings provide support for potential future research that may assess whether such an intervention could positively impact epigenetic aging and other indices of long-term health outcomes. Other studies could also examine different aspects of resilience, such as cultural or environmental factors that contribute to resilience to determine if they also are protective against the effects of stress on epigenetic age acceleration. Future studies could also explore other physiologic mechanisms through which psychological resilience may influence epigenetic aging. Based on prior work, inflammation could be particularly important for this relationship. In particular, prior studies have found C-reactive protein [ 87 ] and IL-6 [ 88 ] to be related to emotion regulation and measures of health. The work by Gianaros et al suggests that neurologic activity of the dorsal anterior cingulate cortex may be involved as well.

The relationship between cumulative stress, epigenetic aging, and insulin resistance is of particular note given the prominence of insulin signaling in aging-related pathways [ 89 , 90 ], as well as current trials investigating metformin as a potential anti-aging drug [ 33 ]. In association with this body of work, our study suggests insulin resistance as at least one factor through which stress is associated with accelerated aging, even in a healthy population not suffering from diabetes. As this study is limited by its cross-sectional nature, any causal hypotheses regarding interactions between stress, BMI, insulin resistance, and aging will require longitudinal data to draw specific inferences beyond correlative relationships. Longitudinal studies would also enable prospective assessments of stress, which may be less subject to recall bias based on their current context. This study also identifies the cortisol/ACTH ratio as a potential point of connection between stress and epigenetic aging. However, this measure is somewhat limited in that it reflects an acute measure of the HPA axis, and this relationship becomes non-significant with the inclusion of our covariates. Future studies could utilize other, longer-term measures of HPA axis function such as hair cortisol to better characterize the relationship between stress, epigenetic aging, and the HPA axis.

Nonetheless, this study is the first to identify a clear relationship between cumulative stress and GrimAge acceleration in a healthy population, which suggests stress may play a role in accelerated aging even prior to the onset of chronic diseases. Notably, this relationship was strongly moderated by resilience factors, including self-control and emotion regulation. We also identified smoking, BMI, insulin signaling, and potentially HPA signaling as mediators of this response. However, even when accounting for all these factors as well as demographic covariates such as race, cumulative stress continues to demonstrate a significant impact on GAA, suggesting other mechanisms relating stress to aging not identified herein are also present.

Code availability

R scripts utilized for data analysis are available by contacting the authors directly.

Roy B, Riley C, Sinha R. Emotion regulation moderates the association between chronic stress and cardiovascular disease risk in humans: a cross-sectional study. Stress. 2018:1-8, https://doi.org/10.1080/10253890.2018.1490724 .

Boehm JK, Kubzansky LD. The heart’s content: the association between positive psychological well-being and cardiovascular health. Psychol Bull. 2012:138;655-691.

Sampasa-Kanyinga H, Chaput J-P. Associations among self-perceived work and life stress, trouble sleeping, physical activity, and body weight among Canadian adults. Preventive Med. 2017;96:16–20. https://doi.org/10.1016/j.ypmed.2016.12.013 .

Article   Google Scholar  

Kelly SJ, Ismail M. Stress and type 2 diabetes: a review of how stress contributes to the development of type 2 diabetes. Annu Rev Public Health. 2015;36:441–62. https://doi.org/10.1146/annurev-publhealth-031914-122921 .

Article   PubMed   Google Scholar  

Liu MY, Li N, Li WA, Khan H. Association between psychosocial stress and hypertension: a systematic review and meta-analysis. Neurol Res. 2017;39:573–80. https://doi.org/10.1080/01616412.2017.1317904 .

Halaris A. Inflammation-associated co-morbidity between depression and cardiovascular disease. Curr Top Behav Neurosci. 2017;31:45–70. https://doi.org/10.1007/7854_2016_28 .

Article   CAS   PubMed   Google Scholar  

Joseph JJ, Golden SH. Cortisol dysregulation: the bidirectional link between stress, depression, and type 2 diabetes mellitus. Ann N. Y Acad Sci. 2017;1391:20–34. https://doi.org/10.1111/nyas.13217 .

Tsounis D, Bouras G, Giannopoulos G, Papadimitriou C, Alexopoulos D, Deftereos S. Inflammation markers in essential hypertension. Med Chem. 2014;10:672–81. https://doi.org/10.2174/1573406410666140318111328 .

Silverman MN, Sternberg EM. Glucocorticoid regulation of inflammation and its functional correlates: from HPA axis to glucocorticoid receptor dysfunction. Ann N. Y Acad Sci. 2012;1261:55–63. https://doi.org/10.1111/j.1749-6632.2012.06633.x .

Article   CAS   PubMed   PubMed Central   Google Scholar  

Miller R, Kirschbaum C. Cultures under stress: A cross-national meta-analysis of cortisol responses to the Trier Social Stress Test and their association with anxiety-related value orientations and internalizing mental disorders. Psychoneuroendocrinology. 2019;105:147–54. https://doi.org/10.1016/j.psyneuen.2018.12.236 .

Giacco D, Laxhman N, Priebe S. Prevalence of and risk factors for mental disorders in refugees. Semin Cell Dev Biol. 2018;77:144–52. https://doi.org/10.1016/j.semcdb.2017.11.030 .

Abravanel BT, Sinha R. Emotion dysregulation mediates the relationship between lifetime cumulative adversity and depressive symptomatology. J Psychiatr Res. 2015;61:89–96. https://doi.org/10.1016/j.jpsychires.2014.11.012 .

Wolff, M, Enge, S, Kräplin, A, Krönke, KM, Bühringer, G, Smolka, MN et al. Chronic stress, executive functioning, and real-life self-control: an experience sampling study. J Pers. 2020, https://doi.org/10.1111/jopy.12587 .

Duckworth AL, Kim B, Tsukayama E. Life stress impairs self-control in early adolescence. Front Psychol. 2012;3:608 https://doi.org/10.3389/fpsyg.2012.00608 .

Lewis EJ, Yoon KL, Joormann J. Emotion regulation and biological stress responding: associations with worry, rumination, and reappraisal. Cogn Emot. 2018;32:1487–98. https://doi.org/10.1080/02699931.2017.1310088 .

Raio CM, Orederu TA, Palazzolo L, Shurick AA, Phelps EA. Cognitive emotion regulation fails the stress test. Proc Natl Acad Sci USA. 2013;110:15139–44. https://doi.org/10.1073/pnas.1305706110 .

Article   PubMed   PubMed Central   Google Scholar  

Sinha R. How does stress increase risk of drug abuse and relapse? Psychopharmacol (Berl). 2001;158:343–59. https://doi.org/10.1007/s002130100917 .

Article   CAS   Google Scholar  

Sinha R. Chronic stress, drug use, and vulnerability to addiction. Ann N. Y Acad Sci. 2008;1141:105–30. https://doi.org/10.1196/annals.1441.030 .

Baumeister RF, Bratslavsky E, Muraven M, Tice DM. Ego depletion: is the active self a limited resource? J Pers Soc Psychol. 1998;74:1252–65. https://doi.org/10.1037//0022-3514.74.5.1252 .

Maier SU, Makwana AB, Hare TA. Acute stress impairs self-control in goal-directed choice by altering multiple functional connections within the brain’s decision circuits. Neuron. 2015;87:621–31. https://doi.org/10.1016/j.neuron.2015.07.005 .

Muraven M, Baumeister RF. Self-regulation and depletion of limited resources: does self-control resemble a muscle? Psychol Bull. 2000;126:247–59. https://doi.org/10.1037/0033-2909.126.2.247 .

Beutel TF, Zwerenz R, Michal M. Psychosocial stress impairs health behavior in patients with mental disorders. BMC Psychiatry. 2018;18:375 https://doi.org/10.1186/s12888-018-1956-8 .

Wemm SE, Sinha R. Drug-induced stress responses and addiction risk and relapse. Neurobiol Stress. 2019;10:100148 https://doi.org/10.1016/j.ynstr.2019.100148 .

Kwarteng JL, Schulz AJ, Mentz GB, Israel BA, Perkins DW. Independent effects of neighborhood poverty and psychosocial stress on obesity over time. J Urban Health. 2017;94:791–802. https://doi.org/10.1007/s11524-017-0193-7 .

Stults-Kolehmainen MA, Sinha R. The effects of stress on physical activity and exercise. Sports Med. 2014;44:81–121. https://doi.org/10.1007/s40279-013-0090-5 .

Chao AM, Jastreboff AM, White MA, Grilo CM, Sinha R. Stress, cortisol, and other appetite-related hormones: Prospective prediction of 6-month changes in food cravings and weight. Obes (Silver Spring). 2017;25:713–20. https://doi.org/10.1002/oby.21790 .

Sinha R, Jastreboff AM. Stress as a common risk factor for obesity and addiction. Biol Psychiatry. 2013;73:827–35. https://doi.org/10.1016/j.biopsych.2013.01.032 .

Sinha R. Role of addiction and stress neurobiology on food intake and obesity. Biol Psychol. 2018;131:5–13. https://doi.org/10.1016/j.biopsycho.2017.05.001 .

Wirtz PH, von Känel R. Psychological stress, inflammation, and coronary heart disease. Curr Cardiol Rep. 2017;19:111 https://doi.org/10.1007/s11886-017-0919-x .

Lavretsky H, Newhouse PA. Stress, inflammation, and aging. Am J Geriatr Psychiatry. 2012;20:729–33. https://doi.org/10.1097/JGP.0b013e31826573cf .

Edes AN, Crews DE. Allostatic load and biological anthropology. Am J Phys Anthropol. 2017;162:44–70. https://doi.org/10.1002/ajpa.23146 . Suppl 63 .

Costantino S, Paneni F, Cosentino F. Ageing, metabolism and cardiovascular disease. J Physiol. 2016;594:2061–73. https://doi.org/10.1113/JP270538 .

Barzilai N, Crandall JP, Kritchevsky SB, Espeland MA. Metformin as a tool to target aging. Cell Metab. 2016;23:1060–5. https://doi.org/10.1016/j.cmet.2016.05.011 .

Mason AE, Hecht FM, Daubenmier JJ, Sbarra DA, Lin J, Moran PJ, et al. Weight loss maintenance and cellular aging in the supporting health through nutrition and exercise study. Psychosom Med. 2018;80:609–19. https://doi.org/10.1097/psy.0000000000000616 .

Puterman E, Lin J, Blackburn E, O’Donovan A, Adler N, Epel E. The power of exercise: buffering the effect of chronic stress on telomere length. PLoS ONE. 2010;5:e10837 https://doi.org/10.1371/journal.pone.0010837 .

Ornish D, Lin J, Chan JM, Epel E, Kemp C, Weidner G, et al. Effect of comprehensive lifestyle changes on telomerase activity and telomere length in men with biopsy-proven low-risk prostate cancer: 5-year follow-up of a descriptive pilot study. Lancet Oncol. 2013;14:1112–20. https://doi.org/10.1016/s1470-2045(13)70366-8 .

Puterman E, Epel ES, Lin J, Blackburn EH, Gross JJ, Whooley MA, et al. Multisystem resiliency moderates the major depression-telomere length association: findings from the Heart and Soul Study. Brain Behav Immun. 2013;33:65–73. https://doi.org/10.1016/j.bbi.2013.05.008 .

Osório C, Probert T, Jones E, Young AH, Robbins I. Adapting to stress: understanding the neurobiology of resilience. Behav Med. 2017;43:307–22. https://doi.org/10.1080/08964289.2016.1170661 .

Sandifer PA, Walker AH. Enhancing disaster resilience by reducing stress-associated health impacts. Front Public Health. 2018;6:373 https://doi.org/10.3389/fpubh.2018.00373 .

Kennedy B, Fang F, Valdimarsdóttir U, Udumyan R, Montgomery S, Fall K. Stress resilience and cancer risk: a nationwide cohort study. J Epidemiol Community Health. 2017;71:947–53. https://doi.org/10.1136/jech-2016-208706 .

Bergh C, Udumyan R, Fall K, Almroth H, Montgomery S. Stress resilience and physical fitness in adolescence and risk of coronary heart disease in middle age. Heart. 2015;101:623–9. https://doi.org/10.1136/heartjnl-2014-306703 .

Bergh C, Udumyan R, Fall K, Nilsagård Y, Appelros P, Montgomery S. Stress resilience in male adolescents and subsequent stroke risk: cohort study. J Neurol Neurosurg Psychiatry. 2014;85:1331–6. https://doi.org/10.1136/jnnp-2013-307485 .

Felix AS, Lehman A, Nolan TS, Sealy-Jefferson S, Breathett K, Hood DB, et al. Stress, resilience, and cardiovascular disease risk among black women. Circ Cardiovasc Qual Outcomes. 2019;12:e005284 https://doi.org/10.1161/circoutcomes.118.005284 .

Mehta D, Bruenig D, Lawford B, Harvey W, Carrillo-Roa T, Morris CP, et al. Accelerated DNA methylation aging and increased resilience in veterans: the biological cost for soldiering on. Neurobiol Stress. 2018;8:112–9. https://doi.org/10.1016/j.ynstr.2018.04.001 .

Boks MP, van Mierlo HC, Rutten BP, Radstake TR, De Witte L, Geuze E, et al. Longitudinal changes of telomere length and epigenetic age related to traumatic stress and post-traumatic stress disorder. Psychoneuroendocrinology. 2015;51:506–12. https://doi.org/10.1016/j.psyneuen.2014.07.011 .

James SA. John Henryism and the health of African-Americans. Cult Med Psychiatry. 1994;18:163–82. https://doi.org/10.1007/BF01379448 .

Gupta S, Belanger E, Phillips SP. Low socioeconomic status but resilient: panacea or double trouble? John Henryism in the International IMIAS Study of Older Adults. J Cross Cult Gerontol. 2019;34:15–24. https://doi.org/10.1007/s10823-018-9362-8 .

Horvath S. DNA methylation age of human tissues and cell types. Genome Biol. 2013;14:R115 https://doi.org/10.1186/gb-2013-14-10-r115 .

Horvath S, Raj K. DNA methylation-based biomarkers and the epigenetic clock theory of ageing. Nat Rev Genet. 2018;19:371–84. https://doi.org/10.1038/s41576-018-0004-3 .

Levine ME, Lu AT, Quach A, Chen BH, Assimes TL, Bandinelli S, et al. An epigenetic biomarker of aging for lifespan and healthspan. Aging (Albany NY). 2018;10:573–91. https://doi.org/10.18632/aging.101414 .

Lu AT, Quach A, Wilson JG, Reiner AP, Aviv A, Raj K, et al. DNA methylation GrimAge strongly predicts lifespan and healthspan. Aging (Albany NY). 2019;11:303–27. https://doi.org/10.18632/aging.101684 .

Bell CG, Lowe R, Adams PD, Baccarelli AA, Beck S, Bell JT, et al. DNA methylation aging clocks: challenges and recommendations. Genome Biol. 2019;20:249 https://doi.org/10.1186/s13059-019-1824-y .

Breitling LP, Saum KU, Perna L, Schöttker B, Holleczek B, Brenner H. Frailty is associated with the epigenetic clock but not with telomere length in a German cohort. Clin Epigenetics. 2016;8:21 https://doi.org/10.1186/s13148-016-0186-5 .

Gao X, Zhang Y, Mons U, Brenner H. Leukocyte telomere length and epigenetic-based mortality risk score: associations with all-cause mortality among older adults. Epigenetics. 2018;13:846–57. https://doi.org/10.1080/15592294.2018.1514853 .

Marioni RE, Harris SE, Shah S, McRae AF, von Zglinicki T, Martin-Ruiz C, et al. The epigenetic clock and telomere length are independently associated with chronological age and mortality. Int J Epidemiol. 2018;45:424–32. https://doi.org/10.1093/ije/dyw041 .

Jylhävä J, Pedersen NL, Hägg S. Biological age predictors. EBioMedicine. 2017;21:29–36. https://doi.org/10.1016/j.ebiom.2017.03.046 .

Fries GR, Bauer IE, Scaini G, Wu MJ, Kazimi IF, Valvassori SS, et al. Accelerated epigenetic aging and mitochondrial DNA copy number in bipolar disorder. Transl Psychiatry. 2017;7:1283 https://doi.org/10.1038/s41398-017-0048-8 .

Palma-Gudiel H, Fananas L, Horvath S, Zannas AS. Psychosocial stress and epigenetic aging. Int Rev Neurobiol. 2020;150:107–28. https://doi.org/10.1016/bs.irn.2019.10.020 .

Zannas AS, Arloth J, Carrillo-Roa T, Iurato S, Röh S, Ressler KJ, et al. Lifetime stress accelerates epigenetic aging in an urban, African American cohort: relevance of glucocorticoid signaling. Genome Biol. 2015;16:266 https://doi.org/10.1186/s13059-015-0828-5 .

Simons RL, Lei MK, Beach SR, Philibert RA, Cutrona CE, Gibbons FX, et al. Economic hardship and biological weathering: The epigenetics of aging in a U.S. sample of black women. Soc Sci Med. 2016;150:192–200. https://doi.org/10.1016/j.socscimed.2015.12.001 .

Chen E, Miller GE, Yu T, Brody GH. The great recession and health risks in African American youth. Brain Behav Immun. 2016;53:234–41. https://doi.org/10.1016/j.bbi.2015.12.015 .

Brody GH, Miller GE, Yu T, Beach SR, Chen E. Supportive family environments ameliorate the link between racial discrimination and epigenetic aging: a replication across two longitudinal cohorts. Psychol Sci. 2016;27:530–41. https://doi.org/10.1177/0956797615626703 .

Wolf EJ, Maniates H, Nugent N, Maihofer AX, Armstrong D, Ratanatharathorn A, et al. Traumatic stress and accelerated DNA methylation age: A meta-analysis. Psychoneuroendocrinology. 2018;92:123–34. https://doi.org/10.1016/j.psyneuen.2017.12.007 .

McCrory C, Fiorito G, Hernandez B, Polidoro S, O’Halloran AM, Hever A, et al. GrimAge outperforms other epigenetic clocks in the prediction of age-related clinical phenotypes and all-cause mortality. J Gerontol A Biol Sci Med Sci. 2020. https://doi.org/10.1093/gerona/glaa286 .

Gratz KL, Roemer L. Multidimensional assessment of emotion regulation and dysregulation: development, factor structure, and initial validation of the difficulties in Emotion Regulation Scale. J Psychopathol Behav Assess. 2004;26:41–54. https://doi.org/10.1023/B:JOBA.0000007455.08539.94 .

Tangney JP, Baumeister RF, Boone AL. High self-control predicts good adjustment, less pathology, better grades, and interpersonal success. J Personal. 2004;72:271–324. https://doi.org/10.1111/j.0022-3506.2004.00263.x .

Xu K, Zhang X, Wang Z, Hu Y, Sinha R. Epigenome-wide association analysis revealed that SOCS3 methylation influences the effect of cumulative stress on obesity. Biol Psychol. 2018;131:63–71. https://doi.org/10.1016/j.biopsycho.2016.11.001 .

Brodman K, Erdmann AJ Jr, Lorge I, Wolff HG, Broadbent TH. The Cornell medical index; a adjunct to medical interview. J Am Med Assoc. 1949;140:530–4. https://doi.org/10.1001/jama.1949.02900410026007 .

Turner RJ, Wheaton B, Lloyd DA. The epidemiology of social stress. Am Sociological Rev. 1995;60:104–25. https://doi.org/10.2307/2096348 .

Ansell EB, Gu P, Tuit K, Sinha R. Effects of cumulative stress and impulsivity on smoking status. Hum Psychopharmacol. 2012;27:200–8. https://doi.org/10.1002/hup.1269 .

Rosseel Y. lavaan: an R package for structural equation modeling. 2012. 2012;48:36 https://doi.org/10.18637/jss.v048.i02 .

Seo D, Tsou KA, Ansell EB, Potenza MN, Sinha R. Cumulative adversity sensitizes neural response to acute stress: association with health symptoms. Neuropsychopharmacology. 2014;39:670–80. https://doi.org/10.1038/npp.2013.250 .

Perlmutter M, Nyquist L. Relationships between self-reported physical and mental health and intelligence performance across adulthood. J Gerontol. 1990;45:P145–155. https://doi.org/10.1093/geronj/45.4.p145 .

Abramson JH. The cornell medical index as an epidemiological tool. Am J Public Health Nations Health. 1966;56:287–98. https://doi.org/10.2105/ajph.56.2.287 .

Revelle W. psych: Procedures for Psychological, Psychometric, and Personality Research, 2021. https://CRAN.R-project.org/package=psych .

Houseman EA, Accomando WP, Koestler DC, Christensen BC, Marsit CJ, Nelson HH, et al. DNA methylation arrays as surrogate measures of cell mixture distribution. BMC Bioinform. 2012;13:86 https://doi.org/10.1186/1471-2105-13-86 .

R Foundation for Statistical Computing. R: a language and environment for statistical computing. R Foundation for Statistical Computing; 2020.

Richardson JTE. Eta squared and partial eta squared as measures of effect size in educational research. Educ Res Rev. 2011;6:135–47. https://doi.org/10.1016/j.edurev.2010.12.001 .

Tingley D, Yamamoto T, Hirose K, Keele L, Imai K. mediation: R package for causal mediation analysis. 2014. 2014;59:38 https://doi.org/10.18637/jss.v059.i05 .

Searle SR, Speed FM, Milliken GA. Population marginal means in the linear model: an alternative to least squares means. Am Statistician. 1980;34:216–21. https://doi.org/10.1080/00031305.1980.10483031 .

Wolf EJ, Morrison FG. Traumatic stress and accelerated cellular aging: from epigenetics to cardiometabolic disease. Curr Psychiatry Rep. 2017;19:75 https://doi.org/10.1007/s11920-017-0823-5 .

Yang, R, Wu, GWY, Verhoeven, JE, Gautam, A, Reus, VI, Kang, JI et al. A DNA methylation clock associated with age-related illnesses and mortality is accelerated in men with combat PTSD. Mol Psychiatry. 2020. https://doi.org/10.1038/s41380-020-0755-z .

Squassina, A, Pisanu, C & Vanni, R mood disorders, accelerated aging, and inflammation: is the link hidden in telomeres? Cells. 2019:8, https://doi.org/10.3390/cells8010052 .

Higgins-Chen AT, Boks MP, Vinkers CH, Kahn RS, Levine ME. Schizophrenia and epigenetic aging biomarkers: increased mortality, reduced cancer risk, and unique clozapine effects. Biol Psychiatry. 2020, https://doi.org/10.1016/j.biopsych.2020.01.025 .

Guendelman S, Medeiros S, Rampes H. Mindfulness and emotion regulation: insights from neurobiological, psychological, and clinical studies. Front Psychol. 2017;8:220–220. https://doi.org/10.3389/fpsyg.2017.00220 .

Roy B, Riley C, Sinha R. Emotion regulation moderates the association between chronic stress and cardiovascular disease risk in humans: a cross-sectional study. Stress (Amst, Neth). 2018;21:548–55. https://doi.org/10.1080/10253890.2018.1490724 .

Appleton AA, Buka SL, Loucks EB, Gilman SE, Kubzansky LD. Divergent associations of adaptive and maladaptive emotion regulation strategies with inflammation. Health Psychol: Off J Div Health Psychol, Am Psychological Assoc. 2013;32:748–56. https://doi.org/10.1037/a0030068 .

Gianaros PJ, Marsland AL, Kuan DCH, Schirda BL, Jennings JR, Sheu LK, et al. An inflammatory pathway links atherosclerotic cardiovascular disease risk to neural activity evoked by the cognitive regulation of emotion. Biol psychiatry. 2014;75:738–45. https://doi.org/10.1016/j.biopsych.2013.10.012 .

Cabreiro F, Au C, Leung KY, Vergara-Irigaray N, Cocheme HM, Noori T, et al. Metformin retards aging in C. elegans by altering microbial folate and methionine metabolism. Cell. 2013;153:228–39. https://doi.org/10.1016/j.cell.2013.02.035 .

Martin-Montalvo A, Mercken EM, Mitchell SJ, Palacios HH, Mote PL, Scheibye-Knudsen M, et al. Metformin improves healthspan and lifespan in mice. Nat Commun. 2013;4:2192 https://doi.org/10.1038/ncomms3192 .

Download references

Acknowledgements

The authors would like to acknowledge the Yale Center of Genome Analysis for DNA methylation profiling. Funding for this study is from NIH Common Fund UL1-DE019586 (R.S.), PL1-DA24859 (R.S.), R01-AA013892 (R.S.), NIH R01DA047063 (K.X.), NIH T32MH019961 (Z.M.H.), NIH R25MH071584 (Z.M.H.). These data were presented at the SOBP virtual conference in April 2021 as a poster.

Author information

Authors and affiliations.

Department of Psychiatry, Yale University, New Haven, CT, USA

Zachary M. Harvanek, Ke Xu & Rajita Sinha

Yale Stress Center, Yale University, New Haven, CT, USA

Nia Fogelman & Rajita Sinha

Department of Psychiatry, Connecticut Veteran Healthcare System, West Haven, CT, USA

Department of Neuroscience, Yale University, New Haven, CT, USA

Rajita Sinha

Child Study Center, Yale University, New Haven, CT, USA

You can also search for this author in PubMed   Google Scholar

Contributions

Z.M.H., K.X., and R.S. conceptualized the project. Z.M.H. and N.F. performed the data analysis, with recommendations from K.X. and R.S. Z.M.H. produced the figures and tables. Z.M.H. wrote the manuscript, and all authors contributed to and edited the manuscript.

Corresponding author

Correspondence to Rajita Sinha .

Ethics declarations

Competing interests.

The authors declare no competing interests.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary information

Supplementary material, supplementary figure 1, supplementary figure 2, supplementary figure 3, rights and permissions.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/ .

Reprints and permissions

About this article

Cite this article.

Harvanek, Z.M., Fogelman, N., Xu, K. et al. Psychological and biological resilience modulates the effects of stress on epigenetic aging. Transl Psychiatry 11 , 601 (2021). https://doi.org/10.1038/s41398-021-01735-7

Download citation

Received : 28 June 2021

Revised : 31 October 2021

Accepted : 10 November 2021

Published : 27 November 2021

DOI : https://doi.org/10.1038/s41398-021-01735-7

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

This article is cited by

Self-control is associated with health-relevant disparities in buccal dna-methylation measures of biological aging in older adults.

  • Y. E. Willems
  • A. deSteiguer
  • Laurel Raffington

Clinical Epigenetics (2024)

Psychogenic Aging: A Novel Prospect to Integrate Psychobiological Hallmarks of Aging

  • Manuel Faria
  • Michael Snyder

Translational Psychiatry (2024)

Interplay Between Skeletal and Hematopoietic Cells in the Bone Marrow Microenvironment in Homeostasis and Aging

  • Emily R. Quarato
  • Noah A. Salama
  • Laura M. Calvi

Current Osteoporosis Reports (2024)

Childhood adversity, accelerated GrimAge, and associated health consequences

  • Zachary M. Harvanek
  • Anastacia Y. Kudinova
  • Audrey R. Tyrka

Journal of Behavioral Medicine (2024)

Mind body medicine: a modern bio-psycho-social model forty-five years after Engel

  • Gregory Fricchione

BioPsychoSocial Medicine (2023)

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

research studies on stress

  • A-Z Publications

Annual Review of Psychology

Volume 72, 2021, review article, stress and health: a review of psychobiological processes.

  • Daryl B. O'Connor 1 , Julian F. Thayer 2 , and Kavita Vedhara 3
  • View Affiliations Hide Affiliations Affiliations: 1 School of Psychology, University of Leeds, Leeds LS2 9JT, United Kingdom; email: [email protected] 2 Department of Psychological Science, School of Social Ecology, University of California, Irvine, California 92697, USA; email: [email protected] 3 Division of Primary Care, School of Medicine, University of Nottingham, Nottingham NG7 2UH, United Kingdom; email: [email protected]
  • Vol. 72:663-688 (Volume publication date January 2021) https://doi.org/10.1146/annurev-psych-062520-122331
  • First published as a Review in Advance on September 04, 2020
  • Copyright © 2021 by Annual Reviews. All rights reserved

The cumulative science linking stress to negative health outcomes is vast. Stress can affect health directly, through autonomic and neuroendocrine responses, but also indirectly, through changes in health behaviors. In this review, we present a brief overview of ( a ) why we should be interested in stress in the context of health; ( b ) the stress response and allostatic load; ( c ) some of the key biological mechanisms through which stress impacts health, such as by influencing hypothalamic-pituitary-adrenal axis regulation and cortisol dynamics, the autonomic nervous system, and gene expression; and ( d ) evidence of the clinical relevance of stress, exemplified through the risk of infectious diseases. The studies reviewed in this article confirm that stress has an impact on multiple biological systems. Future work ought to consider further the importance of early-life adversity and continue to explore how different biological systems interact in the context of stress and health processes.

Article metrics loading...

Full text loading...

Literature Cited

  • Adam EK , Doane LD , Zinbarg RE , Mineka S , Craske MG , Griffith JW 2010 . Prospective prediction of major depressive disorder from cortisol awakening responses in adolescence. Psychoneuroendocrinology 35 : 921– 31 [Google Scholar]
  • Adam EK , Quinn ME , Tavernier R , McQuillan MT , Dahlke KA , Gilbert KE 2017 . Diurnal cortisol slopes and mental and physical health outcomes: a systematic review and meta-analysis. Psychoneuroendocrinology 83 : 25– 41 The first review and meta-analysis of the relationships between diurnal cortisol and health outcomes. [Google Scholar]
  • al'Absi M , Wittmers LE Jr 2003 . Enhanced adrenocortical responses to stress in hypertension-prone men and women. Ann. Behav. Med. 25 : 25– 33 [Google Scholar]
  • Am. Psychol. Assoc 2019 . Stress in America 2019 Washington, DC: Am. Psychol. Assoc. [Google Scholar]
  • Ansorge S , Schön E. 1987 . Dipeptidyl peptidase IV (DP IV), a functional marker of the T lymphocyte system. Acta Histochem 82 : 41– 46 [Google Scholar]
  • Antoni MH , Lutgendorf SK , Blomberg B , Carver CS , Lechner S et al. 2012 . Cognitive-behavioral stress management reverses anxiety-related leukocyte transcriptional dynamics. Biol. Psychiatry 71 : 366– 72 [Google Scholar]
  • Beauchaine TP , Thayer JF. 2015 . Heart rate variability as a transdiagnostic biomarker of psychopathology. Int. J. Psychophysiol. 98 : 338– 50 [Google Scholar]
  • Bellis MA , Hughes K , Leckenby N , Hardcastle KA , Perkins C , Lowey H 2015 . Measuring mortality and the burden of adult disease associated with adverse childhood experiences in England: a national survey. J. Public Health 37 : 445– 54 [Google Scholar]
  • Benarroch EE. 2008 . The arterial baroreflex: functional organization and involvement in neurologic disease. Neurology 71 : 1733– 38 [Google Scholar]
  • Ben-Dov IZ , Kark JD , Ben-Ishay D , Mekler J , Ben-Arie L , Bursztyn M 2007 . Blunted heart rate dip during sleep and all-cause mortality. Arch. Intern. Med. 167 : 2116– 21 [Google Scholar]
  • Bernard K , Frost A , Bennett CB , Lindhiem O 2017 . Maltreatment and diurnal cortisol regulation: a meta-analysis. Psychoneuroendocrinology 78 : 57– 67 [Google Scholar]
  • Black DS , Cole SW , Irwin MR , Breen E , Cyr NMS et al. 2013 . Yogic meditation reverses NF-κB and IRF-related transcriptome dynamics in leukocytes of family dementia caregivers in a randomized controlled trial. Psychoneuroendocrinology 38 : 348– 55 [Google Scholar]
  • Boggero IA , Hostinar CE , Haak EA , Murphy MLM , Segerstrom S 2017 . Psychosocial functioning and the cortisol awakening response: meta-analysis, P-curve analysis and evaluation of the evidential value of existing studies. Biol. Psychol. 129 : 207– 30 [Google Scholar]
  • Bower JE , Greendale G , Crosswell AD , Garet D , Sternlieb B et al. 2014 . Yoga reduces inflammatory signaling in fatigued breast cancer survivors: a randomized controlled trial. Psychoneuroendocrinology 43 : 20– 29 [Google Scholar]
  • Brosschot JF , Verkuil B , Thayer JF 2016 . The default response to uncertainty and the importance of perceived safety in anxiety and stress: an evolution-theoretical perspective. J. Anxiety Disord. 41 : 22– 34 [Google Scholar]
  • Brosschot JF , Verkuil B , Thayer JF 2017 . Exposed to events that never happen: generalized unsafety, the default stress response, and prolonged autonomic activity. Neurosci. Biobehav. Rev. 74 : 287– 96 [Google Scholar]
  • Brosschot JF , Verkuil B , Thayer JF 2018 . Generalized unsafety theory of stress: unsafe environments and conditions, and the default stress response. Int. J. Environ. Res. Public Health 15 : 464 Comprehensive overview of an important new theory of stress. [Google Scholar]
  • Bunea I , Szentágotai-Tătar A , Miu AC 2017 . Early-life adversity and cortisol response to social stress: a meta-analysis. Transl. Psychiatry 7 : 1274 [Google Scholar]
  • Carpenter LL , Carvalho JP , Tyrka AR , Wier LM , Mello AF et al. 2007 . Decreased adrenocorticotropic hormone and cortisol responses to stress in healthy adults reporting significant childhood maltreatment. Biol. Psychiatry 62 : 1080– 87 [Google Scholar]
  • Carpenter LL , Shattuck TT , Tyrka AR , Geracioti TD , Price LH 2011 . Effect of childhood physical abuse on cortisol stress response. Psychopharmacology 214 : 367– 75 [Google Scholar]
  • Carr CP , Martins CM , Stingel AM , Lemgruber VB , Juruena MF 2013 . The role of early life stress in adult psychiatric disorders: a systematic review according to childhood trauma subtypes. J. Nerv. Ment. Dis. 201 : 1007– 20 [Google Scholar]
  • Carroll D , Ginty AT , Whittaker AC , Lovallo WR , de Rooij SR 2017 . The behavioural, cognitive, and neural corollaries of blunted cardiovascular and cortisol reactions to acute psychological stress. Neurosci. Biobehav. Rev. 77 : 74– 86 [Google Scholar]
  • Chida Y , Steptoe A. 2009 . Cortisol awakening response and psychosocial factors: a systematic review and meta-analysis. Biol. Psychol. 80 : 265– 78 [Google Scholar]
  • Clancy F , Prestwich A , Caperon L , O'Connor DB 2016 . Perseverative cognition and health behaviours: a systematic review and meta-analysis. Front. Hum. Neurosci. 10 : 534 The first paper to extend the perseverative cognition hypothesis to health behaviors. [Google Scholar]
  • Clancy F , Prestwich A , Caperon L , Tsipa A , O'Connor DB 2020 . The association between perseverative cognition and sleep in non-clinical populations: a systematic review and meta-analysis. Health Psychol. Rev. In press. https://doi.org/10.1080/17437199.2019.1700819 [Crossref] [Google Scholar]
  • Clow A , Hucklebridge F , Stalder T , Evans P , Thorn L 2010 . The cortisol awakening response: more than a measure of HPA axis function. Neurosci. Biobehav. Rev. 35 : 97– 103 [Google Scholar]
  • Cohen S. 2005 . Keynote presentation at the Eight International Congress of Behavioral Medicine. Int. J. Behav. Med. 12 : 123– 31 [Google Scholar]
  • Cohen S , Doyle WJ , Skoner DP , Rabin BS , Gwaltney JM 1997 . Social ties and susceptibility to the common cold. JAMA 277 : 1940– 44 [Google Scholar]
  • Cohen S , Frank E , Doyle WJ , Skoner DP , Rabin BS , Gwaltney JM Jr 1998 . Types of stressors that increase susceptibility to the common cold in healthy adults. Health Psychol 17 : 214– 23 [Google Scholar]
  • Cohen S , Gianaros PJ , Manuck SB 2016 . A stage model of stress and disease. Perspect. Psychol. Sci. 11 : 456– 63 [Google Scholar]
  • Cohen S , Janicki-Deverts D , Doyle WJ , Miller GE , Frank E et al. 2012 . Chronic stress, glucocorticoid receptor resistance, inflammation, and disease risk. PNAS 109 : 5995– 99 [Google Scholar]
  • Cohen S , Janicki-Deverts D , Turner RB , Doyle WJ 2015 . Does hugging provide stress-buffering social support? A study of susceptibility to upper respiratory infection and illness. Psychol. Sci. 26 : 135– 47 [Google Scholar]
  • Cohen S , Miller GE , Rabin BS 2001 . Psychological stress and antibody response to immunization: a critical review of the human literature. Psychosom. Med. 63 : 7– 18 [Google Scholar]
  • Cohen S , Tyrrell DA , Smith AP 1991 . Psychological stress and susceptibility to the common cold. N. Engl. J. Med. 325 : 606– 12 The first study to show that increases in psychological stress are associated with increased risk of developing a common cold. [Google Scholar]
  • Cole SW. 2013 . Social regulation of human gene expression: mechanisms and implications for public health. Am. J. Public Health 103 : S84– 92 [Google Scholar]
  • Cole SW. 2019 . The conserved transcriptional response to adversity. Curr. Opin. Behav. Sci. 28 : 31– 37 [Google Scholar]
  • Cole SW , Capitanio JP , Chun K , Arevalo JMG , Ma J , Cacioppo JT 2015 . Myeloid differentiation architecture of leukocyte transcriptome dynamics in perceived social isolation. PNAS 112 : 15142– 47 [Google Scholar]
  • Cole SW , Hawkley LC , Arevalo JMG , Cacioppo JT 2011 . Transcript origin analysis identifies antigen-presenting cells as primary targets of socially regulated gene expression in leukocytes. PNAS 108 : 3080– 85 [Google Scholar]
  • Cole SW , Hawkley LC , Arevalo JMG , Sung CY , Rose RM , Cacioppo JT 2007 . Social regulation of gene expression in human leukocytes. Genome Biol 8 : R189 The first study to show that gene expression is influenced by different levels of loneliness. [Google Scholar]
  • Craske MG , Wolitzky-Taylor KB , Mineka S , Zinbarg R , Waters AM et al. 2012 . Elevated responding to safe conditions as a specific risk factor for anxiety versus depressive disorders: evidence from a longitudinal investigation. J. Abnorm. Psychol. 121 : 315– 24 [Google Scholar]
  • Creswell JD , Irwin MR , Burklund LJ , Lieberman MD , Arevalo JM et al. 2012 . Mindfulness-based stress reduction training reduces loneliness and pro-inflammatory gene expression in older adults: a small randomized controlled trial. Brain. Behav. Immun. 26 : 1095– 101 [Google Scholar]
  • Cruz-Pereira JS , Rea K , Nolan YM , O'Leary OF , Dinan TG , Cryan JF 2020 . Depression's unholy trinity: dysregulated stress, immunity, and the microbiome. Annu. Rev. Psychol. 71 : 49– 78 [Google Scholar]
  • Cuspidi C , Giudici V , Negri F , Sala C 2010 . Nocturnal nondipping and left ventricular hypertrophy in hypertension: an updated review. Exp. Rev. Cardiovasc. Ther. 8 : 781– 92 [Google Scholar]
  • Danese A , Baldwin JR. 2017 . Hidden wounds? Inflammatory links between childhood trauma and psychopathology. Annu. Rev. Psychol. 68 : 517– 44 [Google Scholar]
  • Danese A , McEwen BS. 2012 . Adverse childhood experiences, allostasis, allostatic load and age-related disease. Phys. Behav. 106 : 29– 39 [Google Scholar]
  • Dauphinot V , Gosse P , Kossovsky MP , Schott AM , Rouch I et al. 2010 . Autonomic nervous system activity is independently associated with the risk of shift in the non-dipper blood pressure pattern. Hypertens. Res. 33 : 1032– 37 [Google Scholar]
  • De Rooij SR. 2013 . Blunted cardiovascular and cortisol reactivity to acute psychological stress: a summary of results from the Dutch Famine Birth Cohort Study. Int. J. Psychophysiol. 90 : 21– 27 [Google Scholar]
  • De Vugt ME , Nicolson NA , Aalten P , Lousberg R , Jolle J , Verhey FR 2005 . Behavioral problems in dementia patients and salivary cortisol patterns in caregivers. J. Neuropsychiat. Clin. Neurosci. 17 : 201– 7 [Google Scholar]
  • Dickerson SS , Kemeny ME. 2004 . Acute stressors and cortisol responses: a theoretical integration and synthesis of laboratory research. Psychol. Bull. 130 : 355– 91 [Google Scholar]
  • Edwards KM , Burns VE , Allen LM , McPhee JS , Bosch JA et al. 2007 . Eccentric exercise as an adjuvant to influenza vaccination in humans. Brain Behav. Immun. 21 : 209– 17 [Google Scholar]
  • Edwards KM , Burns VE , Reynolds T , Carroll D , Drayson M , Ring C 2006 . Acute stress exposure prior to influenza vaccination enhances antibody response in women. Brain Behav. Immun. 20 : 159– 68 [Google Scholar]
  • Fallo F , Barzon L , Rabbia F , Navarrini C , Conterno A et al. 2002 . Circadian blood pressure patterns and life stress. Psychother. Psychosom. 71 : 350– 56 [Google Scholar]
  • Fan LB , Blumenthal JA , Hinderliter AL , Sherwood A 2013 . The effect of job strain on nighttime blood pressure dipping among men and women with high blood pressure. Scand. J. Work Environ. Health 39 : 112 [Google Scholar]
  • Fortmann AL , Gallo LC. 2013 . Social support and nocturnal blood pressure dipping: a systematic review. Am. J. Hypertens. 26 : 302– 10 [Google Scholar]
  • Fries E , Dettenborn L , Kirschbaum C 2009 . The cortisol awakening response (CAR): facts and future directions. Int. J. Psychophysiol. 72 : 67– 73 [Google Scholar]
  • Gartland N , O'Connor DB , Lawton R , Bristow M 2014 . Exploring day-to-day dynamics of daily stressor appraisals, physical symptoms and the cortisol awakening response. Psychoneuroendocrinology 50 : 130– 38 [Google Scholar]
  • Gasperin D , Netuveli G , Dias-da-Costa JS , Pattussi MP 2009 . Effect of psychological stress on blood pressure increase: a meta-analysis of cohort studies. Cad. Saude Publica 25 : 715– 26 [Google Scholar]
  • Gerritsen L , Geerlings MI , Beekman AT , Deeg DJ , Penninx BW , Comijs HC 2010 . Early and late life events and salivary cortisol in older persons. Psychol. Med. 40 : 1569– 78 [Google Scholar]
  • Glaser R , Kiecolt-Glaser JK , Bonneau RH , Malarkey W , Kennedy S , Hughes J 1992 . Stress-induced modulation of the immune response to recombinant hepatitis B vaccine. Psychosom. Med. 54 : 22– 29 [Google Scholar]
  • Hall M , Vasko R , Buysse D , Ombao H , Chen Q et al. 2004 . Acute stress affects heart rate variability during sleep. Psychosom. Med. 66 : 56– 62 [Google Scholar]
  • Hamer M , Endrighi R , Venuraju SM , Lahiri A , Steptoe A 2012 . Cortisol responses to mental stress and the progression of coronary artery calcification in healthy men and women. PLOS ONE 7 : e31356 [Google Scholar]
  • Hamer M , O'Donnell K , Lahiri A , Steptoe A 2010 . Salivary cortisol responses to mental stress are associated with coronary artery calcification in healthy men and women. Eur. Heart J. 31 : 424– 29 [Google Scholar]
  • Hamer M , Steptoe A. 2012 . Cortisol responses to mental stress and incident hypertension in healthy men and women. J. Clin. Endocrinol. Metab. 97 : E29– 34 [Google Scholar]
  • Hill DC , Moss RH , Sykes-Muskett B , Conner M , O'Connor DB 2018 . Stress and eating behaviors in children and adolescents: systematic review and meta-analysis. Appetite 123 : 14– 22 [Google Scholar]
  • Hughes K , Bellis MA , Hardcastle KA , Sethi D , Butchart A et al. 2017 . The effect of multiple adverse childhood experiences on health: a systematic review and meta-analysis. Lancet Public Health 2 : e356– 66 [Google Scholar]
  • Jarczok MN , Jarczok M , Mauss D , Koenig J , Li J , Herr RM , Thayer JF 2013 . Autonomic nervous system activity and workplace stressors—a systematic review. Neurosci. Biobehav. Rev 37 : 1810– 23 [Google Scholar]
  • Jarczok MN , Jarczok M , Thayer JF 2020 . Work stress and autonomic nervous system activity. Handbook of Socioeconomic Determinants of Occupational Health ed. T Theorellpp. 625 – 56 Cham, Switz: Springer Int. Publ. [Google Scholar]
  • Jarczok MN , Koenig J , Wittling A , Fischer JE , Thayer JF 2019 . First evaluation or an index of low vagally-mediated heart rate variability as a marker of health risks in human adults: proof of concept. J. Clin. Med. 8 : 1940 [Google Scholar]
  • Julius S. 1995 . The defense reaction: a common denominator of coronary risk and blood pressure in neurogenic hypertension. Clin. Exp. Hypertens. 17 : 375– 86 [Google Scholar]
  • Kagan J. 2016 . An overly permissive extension. Perspect. Psychol. Sci. 11 : 442– 50 [Google Scholar]
  • Kirschbaum C , Pirke K-M , Hellhammer DH 1993 . The “Trier Social Stress Test”—a tool for investigating psychobiological stress responses in a laboratory setting. Neuropsychobiol 28 : 76– 81 One of the most popular techniques to induce stress in the laboratory. [Google Scholar]
  • Kivimäki M , Steptoe A. 2018 . Effects of stress on the development and progression of cardiovascular disease. Nat. Rev. Cardiol. 15 : 215 [Google Scholar]
  • Krantz DS , Manuck SB. 1984 . Acute psychophysiologic reactivity and risk of cardiovascular disease: a review and methodologic critique. Psychol. Bull. 96 : 435– 64 [Google Scholar]
  • Landsbergis PA , Dobson M , Koutsouras G , Schnall P 2013 . Job strain and ambulatory blood pressure: a meta-analysis and systematic review. Am. Public Health 103 : e61– 71 [Google Scholar]
  • Lazarus RS. 1999 . Stress and Emotion: A New Synthesis New York: Springer [Google Scholar]
  • Lob E , Steptoe A. 2019 . Cardiovascular disease and hair cortisol: a novel biomarker of chronic stress. Curr. Cardiol. Rep. 21 : 116 [Google Scholar]
  • Loerbroks A , Schilling O , Haxsen V , Jarczok MN , Thayer JF , Fischer JE 2010 . The fruits of one's labor: Effort-reward imbalance but not job strain is related to heart rate variability across the day in 35–44-year-old workers. J. Psychosom. Res. 69 : 151– 59 [Google Scholar]
  • Lovallo WR. 2013 . Early life adversity reduces stress reactivity and enhances impulsive behavior: implications for health behaviors. Int. J. Psychophysiol. 90 : 8– 16 [Google Scholar]
  • Lovallo WR. 2016 . Stress and Health: Biological and Psychological Interactions Thousand Oaks, CA: SAGE. , 3rd ed.. [Google Scholar]
  • Lovallo WR , Cohoon AJ , Sorocco KH , Vincent AS , Acheson A et al. 2019 . Early‐life adversity and blunted stress reactivity as predictors of alcohol and drug use in persons with COMT (rs4680) Val158Met genotypes. Alcohol. Clin. Exp. Res. 43 : 1519– 27 [Google Scholar]
  • Lovallo WR , Dickensheets SL , Myers DA , Thomas TL , Nixon SJ 2000 . Blunted stress cortisol response in abstinent alcoholic and polysubstance-abusing men. Alcohol. Clin. Exp. Res. 24 : 651– 58 Early study showing that a blunted cortisol response may be a marker of HPA axis dysregulation. [Google Scholar]
  • Lovallo WR , Farag NH , Sorocco KH , Acheson A , Cohoon AJ , Vincent AS 2013 . Early life adversity contributes to impaired cognition and impulsive behaviour: studies from the Oklahoma Family Health Patterns Project. Alcohol. Clin. Exp. Res. 37 : 616– 23 [Google Scholar]
  • Lupien SJ , McEwen BS , Gunnar MR , Heim C 2009 . Effects of stress throughout the lifespan on the brain, behaviour and cognition. Nat. Rev. Neurosci. 10 : 434– 45 [Google Scholar]
  • Manuck SB , Krantz DS. 1984 . Psychophysiologic reactivity in coronary heart disease. Behav. Med. Update 6 : 11– 15 [Google Scholar]
  • Marmot M , Brunner E. 2005 . Cohort profile: the Whitehall II study. Int. J. Epidemiol. 34 : 251– 56 [Google Scholar]
  • Marsland AL , Cohen S , Rabin BS , Manuck SB 2006 . Trait positive affect and antibody response to hepatitis B vaccination. Brain Behav. Immun. 20 : 261– 69 [Google Scholar]
  • Massey AJ , Campbell BK , Raine-Fenning N , Pincott-Allen C , Perry J , Vedhara K 2016 . Relationship between hair and salivary cortisol and pregnancy in women undergoing IVF. Psychoneuroendocrinology 74 : 397– 405 [Google Scholar]
  • Matthews K , Schwartz J , Cohen S , Seeman T 2006 . Diurnal cortisol decline is related to coronary calcification: CARDIA study. Psychosom. Med. 68 : 657– 61 [Google Scholar]
  • Mayne SL , Moore KA , Powell-Wiley TM , Evenson KR , Block R , Kershaw KN 2018 . Longitudinal associations of neighborhood crime and perceived safety with blood pressure: the Multi-Ethnic Study of Atherosclerosis (MESA). Am. J. Hypertens. 31 : 1024– 32 [Google Scholar]
  • McEwen BS. 1998 . Protective and damaging effects of stress mediators. N. Engl. J. Med. 338 : 171– 79 Very influential overview of allostasis and allostatic load. [Google Scholar]
  • McEwen BS. 2000 . Allostasis and allostatic load: implications for neuropsychopharmacology. Neuropsychopharmacology 22 : 108– 24 [Google Scholar]
  • McEwen BS. 2018 . Redefining neuroendocrinology: epigenetics of brain-body communication over the life course. Front. Neuroendocrinol. 49 : 8– 30 [Google Scholar]
  • McEwen BS. 2019 . What is the confusion with cortisol. Chronic Stress 3 : 1– 3 [Google Scholar]
  • McEwen BS , McEwen CA. 2016 . Response to Jerome Kagan's Essay on Stress (2016). Perspect. Psychol. Sci. 11 : 451– 55 [Google Scholar]
  • McEwen BS , Seeman T. 1999 . Protective and damaging effects of mediators of stress: elaborating and testing the concepts of allostasis and allostatic load. Ann. N.Y. Acad. Sci. 896 : 30– 47 [Google Scholar]
  • Mehl MR , Raison CL , Pace TW , Arevalo JM , Cole SW 2017 . Natural language indicators of differential gene regulation in the human immune system. PNAS 114 : 12554– 59 [Google Scholar]
  • Miller GE , Chen E. 2006 . Life stress and diminished expression of genes encoding glucocorticoid receptor and β2-adrenergic receptor in children with asthma. PNAS 103 : 5496– 501 [Google Scholar]
  • Miller GE , Chen E , Sze J , Marin T , Arevalo JM et al. 2008 . A functional genomic fingerprint of chronic stress in humans: blunted glucocorticoid and increased NF-κB signaling. Biol. Psychiatry 64 : 266– 72 [Google Scholar]
  • Miller GE , Chen E , Zhou ES 2007 . If it goes up, must it come down? Chronic stress and the hypothalamic-pituitary-adrenocortical axis in humans. Psychol. Bull. 133 : 25– 45 [Google Scholar]
  • Miller GE , Cohen S , Ritchey AK 2002 . Chronic psychological stress and the regulation of pro-inflammatory cytokines: a glucocorticoid-resistance model. Health Psychol 21 : 531 [Google Scholar]
  • Miller GE , Murphy ML , Cashman R , Ma R , Ma J et al. 2014 . Greater inflammatory activity and blunted glucocorticoid signaling in monocytes of chronically stressed caregivers. Brain Behav. Immun. 41 : 191– 99 [Google Scholar]
  • Mortensen J , Dich N , Clark AJ , Ramlau-Hansen C , Head J et al. 2019 . Informal caregiving and diurnal patterns of salivary cortisol: results from the Whitehall II cohort study. Psychoneuroendocrinology 100 : 41– 47 [Google Scholar]
  • Nater UM , Youngblood LS , Jones JF , Unger ER , Miller AH et al. 2008 . Alterations in diurnal salivary cortisol rhythm in a population-based sample of cases with chronic fatigue syndrome. Psychosom. Med. 70 : 298– 305 [Google Scholar]
  • Newman E , O'Connor DB , Conner M 2007 . Daily hassles and eating behaviour: the role of cortisol reactivity. Psychoneuroendocrinology 32 : 125– 32 [Google Scholar]
  • Obrist PA. 1981 . Cardiovascular Psychophysiology: A Perspective New York: Plenum Press [Google Scholar]
  • O'Connor DB , Archer J , Hair WM , Wu FCW 2001 . Activational effects of testosterone on cognitive function in men. Neuropsychologia 39 : 1385– 94 [Google Scholar]
  • O'Connor DB , Branley-Bell D , Green J , Ferguson E , O'Carroll R , O'Connor RC 2020a . Effects of childhood trauma, daily stress and emotions on daily cortisol levels in individuals vulnerable to suicide. J. Abnorm. Psychol. 129 : 92– 107 [Google Scholar]
  • O'Connor DB , Ferguson E , Green J , O'Carroll RE , O'Connor RC 2016 . Cortisol and suicidal behavior: a meta-analysis. Psychoneuroendocrinology 63 : 370– 79 [Google Scholar]
  • O'Connor DB , Gartland N , O'Connor RC 2020b . Stress, cortisol and suicide risk. Int. Rev. Neurobiol. 152 : 101– 30 [Google Scholar]
  • O'Connor DB , Green J , Ferguson E , O'Carroll RE , O'Connor RC 2017 . Cortisol reactivity and suicidal behavior: investigating the role of hypothalamic-pituitary-adrenal (HPA) axis responses to stress in suicide attempters and ideations. Psychoneuroendocrinology 75 : 183– 91 [Google Scholar]
  • O'Connor DB , Green J , Ferguson E , O'Carroll RE , O'Connor RC 2018 . Effects of childhood trauma on cortisol levels in suicide attempters and ideations. Psychoneuroendocrinology 88 : 9– 16 [Google Scholar]
  • O'Connor DB , Hendrickx H , Dadd T , Talbot D , Mayes A et al. 2009 . Cortisol awakening rise in middle-aged women in relation to chronic psychological stress. Psychoneuroendocrinology 34 : 1486– 94 [Google Scholar]
  • O'Connor DB , Jones F , Conner M , McMillan B , Ferguson E 2008 . Effects of daily hassles and eating style on eating behavior. Health Psychol 27 : S20– 31 [Google Scholar]
  • O'Connor DB , Walker S , Hendrickx H , Talbot D , Schaefer A 2013 . Stress-related thinking predicts the cortisol awakening response and somatic symptoms in healthy adults. Psychoneuroendocrinology 38 : 438– 46 [Google Scholar]
  • Ottaviani C , Thayer JF , Verkuil B , Lonigro A , Medea B et al. 2015 . Physiological concomitants of perseverative cognition: a systematic review and meta-analysis. Psychol. Bull. 142 : 231– 59 [Google Scholar]
  • Padden C , Concialdi-McGlynn C , Lydon S 2019 . Psychophysiological measures of stress in caregivers of individuals with autism spectrum disorder: a system review. Dev. Neurorehabilit. 22 : 149– 63 [Google Scholar]
  • Pakulak E , Stevens C , Neville H 2018 . Neuro-, cardio-, and immunoplasticity: effects of early adversity. Annu. Rev. Psychol. 69 : 131– 56 [Google Scholar]
  • Pedersen AF , Zachariae R , Bovbjerg DH 2009 . Psychological stress and antibody response to influenza vaccination: a meta-analysis. Brain Behav. Immun. 23 : 427– 33 [Google Scholar]
  • Plotkin SA. 2010 . Correlates of protection induced by vaccination. Clin. Vaccine Immunol. 17 : 1055– 65 [Google Scholar]
  • Powell DJ , Schlotz W. 2012 . Daily life stress and the cortisol awakening response: testing the anticipation hypothesis. PLOS ONE 7 : e52067 [Google Scholar]
  • Power C , Thomas C , Li L , Hertzman C 2012 . Childhood psychosocial adversity and adult cortisol patterns. Br. J. Psychiatry 201 : 199– 206 [Google Scholar]
  • Pruessner JC , Wolf OT , Hellhammer DH , Buske Kirschbaum A , von Auer K et al. 1997 . Free cortisol levels after awakening: a reliable biological marker for the assessment of adrenocortical activity. Life Sci 61 : 2539– 49 [Google Scholar]
  • Raison CL , Miller AH. 2003 . When not enough is too much: the role of insufficient glucocorticoid signaling in the pathophysiology of stress-related disorders. Am. J. Psychiatry 160 : 1554– 65 [Google Scholar]
  • Raul J-S , Cirimele V , Ludes B , Kintz P 2004 . Detection of physiological concentrations of cortisol and cortisone in human hair. Clin. Biochem. 37 : 1105– 11 [Google Scholar]
  • Roseboom T , de Rooij S , Painter R 2006 . The Dutch famine and its long-term consequences for adult health. Early Hum. Dev. 82 : 485– 91 [Google Scholar]
  • Ruttle PL , Javaras KN , Klein MH , Armstrong JM , Burk LR , Essex MJ 2013 . Concurrent and longitudinal associations between diurnal cortisol and body mass index across adolescence. J. Adolesc. Health 52 : 731– 37 [Google Scholar]
  • Salles GF , Ribeiro FM , Guimarães GM , Muxfeldt ES , Cardoso CR 2014 . A reduced heart rate variability is independently associated with a blunted nocturnal blood pressure fall in patients with resistant hypertension. J. Hypertens. 32 : 644– 51 [Google Scholar]
  • Sapolsky RM , Romero LM , Munck AU 2000 . How do glucocorticoids influence stress responses? Integrating permissive, suppressive, stimulatory, and preparative actions. Endocr. Rev. 21 : 55– 89 [Google Scholar]
  • Schmidt-Reinwald A , Pruessner JC , Hellhammer DH et al. 1999 . The cortisol response to awakening in relation to different challenge tests and a 12-hour cortisol rhythm. Life Sci 64 : 1653– 60 [Google Scholar]
  • Schrepf A , O'Donnell M , Luo Y , Bradley CS , Kreder K et al. 2014 . Inflammation and inflammatory control in interstitial cystitis/bladder pain syndrome: associations with painful symptoms. Pain 155 : 1755– 61 [Google Scholar]
  • Segerstrom SC , Boggero IA , Smith GT , Sephton SE 2014 . Variability and reliability of diurnal cortisol in younger and older adults: implications for design decisions. Psychoneuroendocrinology 49 : 299– 309 [Google Scholar]
  • Segerstrom SC , Miller G. 2004 . Psychological stress and the human immune system: a meta-analytic study of 30 years of inquiry. Psychol. Bull. 130 : 601– 30 The most comprehensive review of the relationship between stress and the immune system. [Google Scholar]
  • Segerstrom SC , O'Connor DB. 2012 . Stress, health and illness: four challenges for the future. Psychol. Health 27 : 128– 40 [Google Scholar]
  • Selye H. 1936 . A syndrome produced by diverse nocuous agents. Nature 138 : 3479 32 [Google Scholar]
  • Selye H. 1950 . Stress and the general adaptation syndrome. BMJ 1 : 1383– 92 [Google Scholar]
  • Selye H. 1951 . The general-adaptation-syndrome. Annu. Rev. Med. 2 : 327– 42 [Google Scholar]
  • Slavich GM. 2019 . Stressnology: the primitive (and problematic) study of life stress exposure and pressing need for better measurement. Brain Behav. Immun. 75 : 3– 5 [Google Scholar]
  • Snyder-Mackler N , Sanz J , Kohn JN , Brinkworth JF , Morrow S et al. 2016 . Social status alters immune regulation and response to infection in macaques. Science 354 : 6315 1041– 45 [Google Scholar]
  • Stalder T , Kirschbaum C , Kudielka BM , Adam EK , Pruessner JC et al. 2016 . Assessment of the cortisol awakening response: expert consensus guidelines. Psychoneuroendocrinology 63 : 414– 32 [Google Scholar]
  • Stalder T , Steudte-Schmiedgen S , Alexander N , Klucken T , Vater A et al. 2017 . Stress-related and basic determinants of hair cortisol in humans: a meta-analysis. Psychoneuroendocrinology 77 : 261– 74 [Google Scholar]
  • Steptoe A , Hamer M , Lin J , Blackburn EH , Erusalimsky JD 2017 . The longitudinal relationship between cortisol responses to mental stress and leukocyte telomere attrition. J. Clin. Endocrinol. Metab. 102 : 962– 69 [Google Scholar]
  • Steptoe A , Serwinski B. 2016 . Cortisol awakening response. Stress: Concepts, Cognition, Emotion, and Behavior G Fink 277– 83 London: Academic [Google Scholar]
  • Sterling P , Eyer J. 1988 . Allostasis: a new paradigm to explain arousal pathology. Handbook of Life Stress, Cognition and Health S Fisher, J Reason 629– 49 New York: Wiley [Google Scholar]
  • Thayer JF , Lane RD. 2007 . The role of vagal function in the risk for cardiovascular disease and mortality. Biol. Psychol. 74 : 224– 42 [Google Scholar]
  • Thayer JF , Yamamoto SS , Brosschot JF 2010 . The relationship of autonomic imbalance, heart rate variability and cardiovascular disease risk factors. Int. J. Cardiol. 141 : 122– 31 Key review linking autonomic nervous system imbalance to health outcomes. [Google Scholar]
  • Thorn L , Hucklebridge F , Evans P , Clow A 2006 . Suspected nonadherence and weekend versus week day differences in the awakening cortisol response. Psychoneuroendocrinology 31 : 1009– 18 [Google Scholar]
  • Thrasher TN. 2006 . Arterial baroreceptor input contributes to long-term control of blood pressure. Curr. Hypertens. Rep. 8 : 249– 54 [Google Scholar]
  • Tomfohr L , Cooper DC , Mills PJ , Nelesen RA , Dimsdale JE 2010 . Everyday discrimination and nocturnal blood pressure dipping in black and white Americans. Psychosom. Med. 72 : 266 [Google Scholar]
  • Tomiyama AJ. 2019 . Stress and obesity. Annu. Rev. Psychol. 70 : 703– 18 [Google Scholar]
  • Turner-Cobb JM , Rixon L , Jessop DS 2011 . Hypothalamic-pituitary-adrenal axis activity and upper respiratory tract infection in young children transitioning to primary school. Psychopharmacology 214 : 309– 17 [Google Scholar]
  • UK Health Saf. Executive 2019 . Work Related Stress, Anxiety and Depression Statistics in Great Britain 2019 London: Crown [Google Scholar]
  • Vedhara K , Ayling K , Sunger K , Caldwell D , Halliday V et al. 2019 . Psychological interventions as vaccine adjuvants: a systematic review. Vaccine 37 : 3255– 66 [Google Scholar]
  • Vedhara K , Cox NK , Wilcock GK , Perks P , Hunt M et al. 1999a . Chronic stress in elderly carers of dementia patients and antibody response to influenza vaccination. Lancet 353 : 627– 31 [Google Scholar]
  • Vedhara K , Fox J , Wang E 1999b . The measurement of stress-related immune dysfunction in psychoneuroimmunology. Neurosci. Biobehav. Rev. 23 : 699– 715 [Google Scholar]
  • Vedhara K , Tuinstra J , Miles JNV , Sanderman R , Ranchor AV 2006 . Psychosocial factors associated with indices of cortisol production in women with breast cancer and controls. Psychoneuroendocrinology 31 : 299– 311 [Google Scholar]
  • Waehrer GM , Miller TR , Silverio Marques SC , Oh DL , Burke Harris N 2020 . Disease burden of adverse childhood experiences across 14 states. PLOS ONE 15 : e0226134 [Google Scholar]
  • Wright KD , Hickman R , Laudenslager ML 2015 . Hair cortisol analysis: a promising biomarker of HPA activation in older adults. Gerontologist 55 : S140– 45 [Google Scholar]
  • Wust S , Federenko I , Hellhammer DH , Kirschbaum C 2000 . Genetic factors, perceived chronic stress and the free cortisol response to awakening. Psychoneuroendocrinology 25 : 707– 20 [Google Scholar]
  • Article Type: Review Article

Most Read This Month

Most cited most cited rss feed, job burnout, executive functions, social cognitive theory: an agentic perspective, on happiness and human potentials: a review of research on hedonic and eudaimonic well-being, sources of method bias in social science research and recommendations on how to control it, mediation analysis, missing data analysis: making it work in the real world, grounded cognition, personality structure: emergence of the five-factor model, motivational beliefs, values, and goals.

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings
  • My Bibliography
  • Collections
  • Citation manager

Save citation to file

Email citation, add to collections.

  • Create a new collection
  • Add to an existing collection

Add to My Bibliography

Your saved search, create a file for external citation management software, your rss feed.

  • Search in PubMed
  • Search in NLM Catalog
  • Add to Search

Does the perception that stress affects health matter? The association with health and mortality

Affiliation.

  • 1 Department of Population Health Sciences, University of Wisconsin-Madison, Madison, WI 53726, USA.
  • PMID: 22201278
  • PMCID: PMC3374921
  • DOI: 10.1037/a0026743

Objective: This study sought to examine the relationship among the amount of stress, the perception that stress affects health, and health and mortality outcomes in a nationally representative sample of U.S. adults.

Methods: Data from the 1998 National Health Interview Survey were linked to prospective National Death Index mortality data through 2006. Separate logistic regression models were used to examine the factors associated with current health status and psychological distress. Cox proportional hazard models were used to determine the impact of perceiving that stress affects health on all-cause mortality. Each model specifically examined the interaction between the amount of stress and the perception that stress affects health, controlling for sociodemographic, health behavior, and access to health care factors.

Results: 33.7% of nearly 186 million (unweighted n = 28,753) U.S. adults perceived that stress affected their health a lot or to some extent. Both higher levels of reported stress and the perception that stress affects health were independently associated with an increased likelihood of worse health and mental health outcomes. The amount of stress and the perception that stress affects health interacted such that those who reported a lot of stress and that stress impacted their health a lot had a 43% increased risk of premature death (HR = 1.43, 95% CI [1.2, 1.7]).

Conclusions: High amounts of stress and the perception that stress impacts health are each associated with poor health and mental health. Individuals who perceived that stress affects their health and reported a large amount of stress had an increased risk of premature death.

PsycINFO Database Record (c) 2012 APA, all rights reserved.

PubMed Disclaimer

Similar articles

  • The Association Between Perceived Stress and Mortality Among People With Multimorbidity: A Prospective Population-Based Cohort Study. Prior A, Fenger-Grøn M, Larsen KK, Larsen FB, Robinson KM, Nielsen MG, Christensen KS, Mercer SW, Vestergaard M. Prior A, et al. Am J Epidemiol. 2016 Aug 1;184(3):199-210. doi: 10.1093/aje/kwv324. Epub 2016 Jul 11. Am J Epidemiol. 2016. PMID: 27407085
  • Comparison of Health and Health Risk Factors Between Lesbian, Gay, and Bisexual Adults and Heterosexual Adults in the United States: Results From the National Health Interview Survey. Gonzales G, Przedworski J, Henning-Smith C. Gonzales G, et al. JAMA Intern Med. 2016 Sep 1;176(9):1344-51. doi: 10.1001/jamainternmed.2016.3432. JAMA Intern Med. 2016. PMID: 27367843
  • Perceived physical activity and mortality: Evidence from three nationally representative U.S. samples. Zahrt OH, Crum AJ. Zahrt OH, et al. Health Psychol. 2017 Nov;36(11):1017-1025. doi: 10.1037/hea0000531. Epub 2017 Jul 20. Health Psychol. 2017. PMID: 28726475
  • Serious psychological distress, as measured by the K6, and mortality. Pratt LA. Pratt LA. Ann Epidemiol. 2009 Mar;19(3):202-9. doi: 10.1016/j.annepidem.2008.12.005. Ann Epidemiol. 2009. PMID: 19217003
  • Being poor and coping with stress: health behaviors and the risk of death. Krueger PM, Chang VW. Krueger PM, et al. Am J Public Health. 2008 May;98(5):889-96. doi: 10.2105/AJPH.2007.114454. Epub 2008 Apr 1. Am J Public Health. 2008. PMID: 18382003 Free PMC article.
  • The risk and protective factors on the mental health of healthcare workers during the lockdown period due to covid-19 pandemic. Liang K, Yang Y, Chen K, Lv F, Du L. Liang K, et al. Sci Rep. 2024 May 21;14(1):11628. doi: 10.1038/s41598-024-62288-5. Sci Rep. 2024. PMID: 38773200 Free PMC article.
  • Association between adult attachment and mental health states among health care workers: the mediating role of social support. Yang Y, Chen K, Liang K, Du W, Guo J, Du L. Yang Y, et al. Front Psychol. 2024 Mar 7;15:1330581. doi: 10.3389/fpsyg.2024.1330581. eCollection 2024. Front Psychol. 2024. PMID: 38515978 Free PMC article.
  • The missing hallmark of health: psychosocial adaptation. López-Otín C, Kroemer G. López-Otín C, et al. Cell Stress. 2024 Mar 12;8:21-50. doi: 10.15698/cst2024.03.294. eCollection 2024. Cell Stress. 2024. PMID: 38476764 Free PMC article.
  • Associations Between Psychological Factors and Adherence to Health Behaviors After Percutaneous Coronary Intervention: The Role of Cardiac Rehabilitation. Douma ER, Kop WJ, Kupper N. Douma ER, et al. Ann Behav Med. 2024 Apr 11;58(5):328-340. doi: 10.1093/abm/kaae008. Ann Behav Med. 2024. PMID: 38431284 Free PMC article.
  • Bedtime negative affect, sleep quality and subjective health in rural China. Sun J, Zhang N, Carter J, Vanhoutte B, Wang J, Chandola T. Sun J, et al. BMC Public Health. 2024 Jan 23;24(1):280. doi: 10.1186/s12889-024-17779-5. BMC Public Health. 2024. PMID: 38263032 Free PMC article.
  • APA. Stress in America. American Psychological Association; 2008.
  • Braveman PA, Egerter SA, Mockenhaupt RE. Broadening the focus: the need to address the social determinants of health. American Journal of Preventive Medicine. 2011;40(1 Suppl 1):S4–18. - PubMed
  • Burazeri G, Goda A, Sulo G, Stefa J, Kark JD. Financial loss in pyramid savings schemes, downward social mobility and acute coronary syndrome in transitional Albania. Journal of Epidemiology and Community Health. 2008;62(7):620–626. - PubMed
  • CDC. Deaths, Percent of Total Deaths, and Death Rates for the 15 Leading Causes of Death: United States and Each State, 1999–2007. Center for Disease Control/National Center for Health Statistics; 2011.
  • Duncan GE, Sydeman SJ, Perri MG, Limacher MC, Martin AD. Can sedentary adults accurately recall the intensity of their physical activity? Preventive Medicine. 2001;33(1):18–26. - PubMed

Publication types

  • Search in MeSH

Grants and funding

  • P30 HD003352/HD/NICHD NIH HHS/United States
  • T32 AG000129-21/AG/NIA NIH HHS/United States
  • T32 AG000129/AG/NIA NIH HHS/United States
  • T32 HS00083/HS/AHRQ HHS/United States
  • T32 HS000083-12/HS/AHRQ HHS/United States
  • T32 HS000083/HS/AHRQ HHS/United States
  • HS000083/HS/AHRQ HHS/United States
  • HS000083-12/HS/AHRQ HHS/United States
  • T32 AG00129/AG/NIA NIH HHS/United States
  • P30 AG17266/AG/NIA NIH HHS/United States
  • R24 HD047873/HD/NICHD NIH HHS/United States

LinkOut - more resources

Full text sources.

  • American Psychological Association
  • Europe PubMed Central
  • Ovid Technologies, Inc.
  • PubMed Central
  • MedlinePlus Health Information
  • Citation Manager

NCBI Literature Resources

MeSH PMC Bookshelf Disclaimer

The PubMed wordmark and PubMed logo are registered trademarks of the U.S. Department of Health and Human Services (HHS). Unauthorized use of these marks is strictly prohibited.

The Social Psychology of Stress, Health, and Coping

  • First Online: 01 January 2013

Cite this chapter

research studies on stress

  • Deborah Carr Ph.D. 4 &
  • Debra Umberson Ph.D. 5  

Part of the book series: Handbooks of Sociology and Social Research ((HSSR))

16k Accesses

34 Citations

3 Altmetric

The study of stress and health is one of the richest areas of research in the social and biomedical sciences. In this chapter, we first describe core concepts in the study of stress, coping, and health. Second, we summarize key theoretical perspectives that frame social psychological research on stress and health. Third, we review the methods and measures used, as well as limitations associated with these approaches. We draw on examples of empirical studies exploring stressors across multiple life domains, including early life adversity, work, family, and environmental strains, and show their impact on a range of physical and mental health outcomes. We also highlight gender, race, SES, and life course differences regarding the prevalence and nature of stress, coping resources, and stress outcomes. We conclude by suggesting directions for future research on stress, health and coping.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Subscribe and save.

  • Get 10 units per month
  • Download Article/Chapter or eBook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
  • Available as EPUB and PDF
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
  • Durable hardcover edition

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Similar content being viewed by others

research studies on stress

The Stress Process: Its Origins, Evolution, and Future

research studies on stress

Stress and Emotions

research studies on stress

Stress Biomarkers as an Objective Window on Experience

Almeida, D. M. (2005). Resilience and vulnerability to daily stressors assessed via diary methods. Current Directions in Psychological Science, 14 , 62–68.

Google Scholar  

Amato, P. (2000). The consequences of divorce for adults and children. Journal of Marriage and Family, 62 , 1269–1287.

Aneshensel, C. S., Rutter, C. M., & Lachenbruch, P. A. (1991). Social structure, stress, and mental health: Competing conceptual and analytic models. American Sociological Review, 56 , 166–178.

Antonucci, T. C. (1990). Social supports and social relationships. In R. H. Binstock & L. K. George (Eds.), The handbook of aging and the social sciences (3rd ed., pp. 205–226). San Diego, CA: Academic.

Avison, W. R., & Turner, R. J. (1988). Stressful life events and depressive symptoms: Disaggregating the effects of acute stressors and chronic strains. Journal of Health and Social Behavior, 29 , 253–264.

Axinn, W., & Pearce, L. D. (2006). Mixed method data collection strategies . New York: Cambridge University Press.

Bandura, A. (1977). Self-efficacy: The exercise of control . New York: Freeman.

Biddle, B. J. (1979). Role theory: Expectations, identities, and behaviors . New York: Academic.

Billings, A. G., & Moos, R. H. (1981). The role of coping resources in attenuating the stress of life events. Journal of Behavioral Medicine, 7 , 139–157.

Bolger, N., DeLongis, A., Kessler, R. C., & Wethington, E. (1989). The contagion of stress across multiple roles. Journal of Marriage and Family, 51 , 175–183.

Breslau, J., Aguilar-Gaxiola, S., Kendler, K. S., Su, M., Williams, D., & Kessler, R. C. (2006). Specifying race-ethnic differences in risk for psychiatric disorder in a USA national sample. Psychologie Médicale, 36 , 57–68.

Brody, L. R., & Hall, J. A. (2010). Gender, emotion, and socialization. In J. C. Chrisler & D. R. McCready (Eds.), Handbook of gender research in psychology . New York: Springer.

Brown, T. (2003). Critical race theory speaks to the sociology of mental health: Mental health problems produced by racial stratification. Journal of Health and Social Behavior, 44 , 292–301.

Burgard, S. A., & Ailshire, J. A. (2009). Putting work to bed: Stressful experiences on the job and sleep quality. Journal of Health and Social Behavior, 50 , 476–492.

Burgard, S., Brand, J., & House, J. S. (2009). Perceived job insecurity and worker health in the United States. Social Science & Medicine, 69 , 777–785.

Campbell, J. C. (2002). Health consequences of intimate partner violence. Lancet, 35 , 1331–1337.

Carlson, D. L. (2010). Well, what did you expect?: Family transitions, life course expectations, and mental health . Doctoral dissertation, The Ohio State University, Columbus, OH.

Carr, D. (2002). The psychological consequences of work-family tradeoffs for three cohorts of women and men. Social Psychology Quarterly, 65 , 103–124.

Carr, D. (2004). Gender, pre-loss marital dependence and older adults’ adjustment to widowhood. Journal of Marriage and Family, 66 , 220–235.

Carr, D., & Springer, K. W. (2010). Advances in families and health research in the 21st century. Journal of Marriage and Family, 72 , 744–761.

Carstensen, L. L., & Turk-Charles, S. (1994). The salience of emotion across the adult life span. Psychology and Aging, 9 , 259–264.

Carver, C. S., Scheier, M. F., & Weintraub, J. K. (1989). Assessing coping strategies: A theoretically based approach. Journal of Personality and Social Psychology, 56 , 267–283.

Caspi, A., Sugden, K., Moffitt, T. E., Taylor, A., Craig, I. W., Harrington, H., et al. (2003). Influence of life stress on depression: Moderation by a polymorphism in the 5-HTT gene. Science, 301 , 386–389.

Caspi, A., Wright, B. R., Moffitt, T. E., & Silva, P. A. (1998). Early failure in the labor market: Childhood and adolescent predictors of unemployment in the transition to adulthood. American Sociological Review, 63 , 424–451.

Cohen, S., Janicki-Deverts, D., & Miller, G. E. (2007). Psychological stress and disease. Journal of the American Medical Association, 298 , 1685–1687.

Cohen, S., & Wills, T. A. (1985). Stress, social support, and the buffering hypothesis. Psychological Bulletin, 109 , 5–24.

Dannefer, D. (2003). Cumulative advantage/disadvantage and the life course: Cross-fertilizing age and social science theory. Journal of Gerontology, 58B , S327–S337.

Downey, L., & Van Willigen, M. (2005). Environmental stressors: The mental health impact of living near industrial activity. Journal of Health and Social Behavior, 46 , 289–305.

Duffy, S. (2000). Ethical and social issues in incorporating genetic research into survey studies. In C. E. Finch, J. W. Vaupel, & K. Kinsella (Eds.), Cells and surveys: Should biological measures be included in social science research? (pp. 303–328). Washington, DC: National Academy Press.

Duncan, G. J., Ziol-Guest, K. M., & Kalil, A. (2010). Early childhood poverty and adult attainment, behavior, and health. Child Development, 81 , 306–325.

Elder, G. H., Jr. (1995). The life course paradigm: Social change and individual development. In P. Moen, G. H. Elder Jr., & K. Liischer (Eds.), Examining lives in context: Perspectives on the ecology of human development (pp. 101–139). Washington, DC: American Psychological Association.

Entringer, S. E., Epel, S. E., Kumsta, R., Lin, J., Hellhammer, D. H., Blackburn, E., et al. (2011). Stress exposure in intrauterine life is associated with shorter telomere length in young adulthood. PNAS, 108 , E513.

Finch, B. K., Hummer, R. A., Kol, B., & Vega, W. A. (2001). The role of discrimination and acculturative stress in the physical health of Mexican origin adults. Hispanic Journal of Behavioral Sciences, 23 , 399–429.

Friedman, E. M., Williams, D. R., Singer, B. H., & Ryff, C. D. (2009). Chronic discrimination predicts higher circulating levels of E-selectin in a national sample: The MIDUS study. Brain, Behavior, and Immunity, 23 , 684–692.

Frone, M. R. (2000). Work–family conflict and employee psychiatric disorders: The National Comorbidity Survey. Journal of Applied Psychology, 85 , 888–895.

Galvan, F. H., & Caetano, R. (2003). Alcohol use and related problems among ethnic minorities in the United States. Alcohol Research & Health, 27 , 87–94.

George, L. K. (1999). Life-course perspectives on mental health. In C. Aneshensel & J. Phelan (Eds.), Handbook of the sociology of mental health (pp. 565–583). New York: Kluwer.

George, L. K., & Lynch, S. M. (2003). Race differences in depressive symptoms: A dynamic perspective on stress exposure and vulnerability. Journal of Health and Social Behavior, 44 , 353–369.

Goldman, N. (1994). Social factors and health: The causation-selection issue revisited. Proceedings of the National Academy of Sciences of the United States of America, 91 , 1251–1255.

Gotlib, I., & Wheaton, B. (Eds.). (1997). Stress and adversity over the life course: Trajectories and turning points . New York: Cambridge University Press.

Gove, W., & Tudor, J. (1973). Adult sex roles and mental illness. The American Journal of Sociology, 77 , 812–835.

Greenfield, E. A., & Marks, N. F. (2006). Linked lives: Adult children’s problems and their parents’ psychological and relational well-being. Journal of Marriage & Family, 68 , 442–454.

Grzywacz, J. G., Almeida, D. M., & McDonald, D. A. (2002). Work-family spillover and daily reports of work and family stress in the adult labor force. Family Relations, 51 , 28–36.

Haas, S. A. (2008). Trajectories of functional health: The ‘long arm’ of childhood health and socioeconomic factors. Social Science & Medicine, 66 , 849–861.

Heckman, J., & Singer, B. (1984). A method for minimizing the impact of distributional assumptions in econometric models for duration data. Econometrica, 52 , 271–320.

Heffner, K. L., Loving, T. J., Kiecolt-Glaser, J. K., Himawan, L. K., Glaser, R., & Malarkey, W. B. (2006). Older spouses’ cortisol responses to marital conflict: Associations with demand/withdraw communication patterns. Journal of Behavioral Medicine, 29 , 317–325.

Holmes, T. A., & Rahe, R. H. (1967). The social readjustment rating scale. Journal of Psychosomatic Research, 11 , 213–218.

Horwitz, A. V. (2002). Outcomes in the sociology of mental health and illness: Where have we been and where are we going. Journal of Health and Social Behavior, 43 , 143–151.

Horwitz, A. V., White, H. R., & Howell-White, S. (1996). The use of multiple outcomes in stress research: A case study of gender differences in responses to marital dissolution. Journal of Health and Social Behavior, 37 , 273–291.

House, J. S. (1977). The three faces of social psychology. Sociometry, 40 , 161–177.

House, J. S., & Kahn, R. L. (1985). Measures and concepts of social support. In S. Cohen & S. L. Syme (Eds.), Social support and health (pp. 83–108). Orlando, FL: Academic.

House, J. S., & Williams, D. R. (2000). Understanding and reducing socioeconomic and racial/ethnic disparities in health. In B. D. Smedley & S. L. Syme (Eds.), Promoting health: Intervention strategies from social and behavior research (pp. 81–124). Washington, DC: National Academy Press.

Idler, E., Boulifard, D., & Contrada, R. (2012). Mending broken hearts: Marriage and mortality following cardiac surgery. Journal of Health and Social Behavior, 53 , 33–49.

Jackson, J. S., Knight, K. M., & Rafferty, J. A. (2010). Race and unhealthy behaviors: Chronic stress, the HPA axis, and physical and mental health disparities over the life course. American Journal of Public Health, 100 , 933–939.

Jackson, P. B. (1997). Role occupancy and minority mental health. Journal of Health and Social Behavior, 38 , 237–255.

Kahn, J., & Pearlin, L. (2006). The impact of financial strain on health over the life course. Journal of Health and Social Behavior, 47 , 17–31.

Kahneman, D., Krueger, A., Schkade, D., Schwarz, M., & Stone, A. (2004). A survey method for characterizing daily life experience: The Day Reconstruction Method (DRM). Science, 306 , 1176–1780.

Kasper, J. D., Ensminger, M. E., Green, K. M., Fothergill, K. E., Juon, H. S., Robertson, J., et al. (2008). Effects of poverty and family stress over three decades on the functional status of older African American women. Journal of Gerontology: Social Sciences, 63B , S201–S210.

Kassel, J. D., Stroud, L. R., & Paronis, C. A. (2003). Smoking, stress, and negative affect: Correlation, causation, and context across stages of smoking. Psychological Bulletin, 129 , 270–304.

Kawachi, I., & Berkman, L. F. (2001). Social ties and mental health. Journal of Urban Health: Bulletin of the New York Academy of Medicine, 78 , 458–467.

Kessler, R. C. (2002). The categorical versus dimensional assessment controversy in the sociology of mental illness. Journal of Health and Social Behavior, 43 , 171–188.

Kessler, R. C., McGonagle, K. A., Zhao, S., Nelson, C. B., Hughes, M., Eshleman, S., et al. (1994). Lifetime and 12-month prevalence of DSM-III-R psychiatric disorders in the United States: Results from the National Comorbidity Survey. Archives of General Psychiatry, 51 , 8–19.

Kessler, R. C., Mickelson, K. D., & Williams, D. (1999). The prevalence, distribution, and mental health correlates of perceived discrimination in the United States. Journal of Health and Social Behavior, 40 , 208–230.

Krause, N., & Borawski-Clark, E. (1995). Social class differences in social support among older adults. The Gerontologist, 35 , 498–508.

Krieger, N., Chen, J. T., Waterman, P. D., Soobader, M. J., Subramanian, S. V., & Carson, R. (2002). Geocoding and monitoring of US socioeconomic inequalities in mortality and cancer incidence: Does the choice of area-based measure and geographic level matter? American Journal of Epidemiology, 156 , 471–482.

Kristenson, H. R., Eriksen, J. K., Sluiter, D., & Starke, H. U. (2004). Psychobiological mechanisms of socio-economic differences in health. Social Science & Medicine, 58 , 1511–1522.

Lawrence, J., Ashford, K., & Dent, P. (2006). Gender differences in coping strategies of undergraduate students and their impact on self-esteem and attainment. Active Learning in Higher Education, 7 , 273–281.

Lazarus, R. S., & Folkman, S. (1984). Stress, appraisal, and coping . New York: Springer.

Lee, S., Colditz, G., Berkman, L., & Kawachi, I. (2003). Caregiving and risk of coronary heart disease in U.S. women: A prospective study. American Journal of Preventive Medicine, 24 , 113–119.

Little, R. J., & Schenker, N. (1995). Missing data. In G. Arminger, C. C. Clogg, & M. E. Sobel (Eds.), Handbook of statistical modeling for the social and behavioral sciences (pp. 39–75). New York: Plenum.

Massoglia, M. (2008). Incarceration as exposure: The prison, infectious disease, and other stress-related illnesses. Journal of Health and Social Behavior, 49 , 56–71.

McEwen, B. S. (1998). Stress, adaptation, and disease: Allostasis and allostatic load. Annals of the New York Academy of Science, 840 , 33–44.

Menaghan, E. G. (1983). Individual coping efforts: Moderators of the relationship between life stress and mental health outcomes. In H. B. Kaplan (Ed.), Psychosocial stress: Trends in theory and research (pp. 157–191). New York: Academic.

Merton, R. K. (1968). The Matthew effect in science: The reward and communication system of science. Science, 199 , 55–63.

Meyer, I. H., Schwartz, S., & Frost, D. M. (2008). Social patterning of stress and coping: Does disadvantaged status confer excess exposure and fewer coping resources? Social Science & Medicine, 67 , 368–379.

Mirowsky, J., & Ross, C. E. (2002a). Measurement for a human science. Journal of Health and Social Behavior, 43 , 152–170.

Mirowsky, J., & Ross, C. E. (2002b). Depression, parenthood, and age at first birth. Social Science & Medicine, 54 , 1281–1298.

Montez, J. K., & Hayward, M. D. (2011). Early life conditions and later life mortality. In R. G. Rogers & E. Crimmins (Eds.), International handbook of adult mortality (pp. 187–206). New York: Springer.

Mossakowski, K. (2003). Coping with perceived discrimination: Does ethnic identity protect mental health? Journal of Health and Social Behavior, 44 , 318–331.

Ng, D. M., & Jeffery, R. W. (2003). Relationships between perceived stress and health behaviors in a sample of working adults. Health Psychology, 22 , 638–642.

Nolen-Hoeksema, S., & Ahrens, C. (2001). Age differences and similarities in the correlates of depressive symptoms. Psychology and Aging, 17 , 116–124.

O’Donovan, A., Tomiyama, A. J., Lin, J., Puterman, E., Adler, N. E., Kemeny, M., et al. (2012). Stress appraisals and cellular aging: A key role for anticipatory threat in the relationship between psychological stress and telomere length. Brain, Behavior, and Immunity, 28 , 573–579.

Parkes, C. M. (1965). Bereavement and mental illness: Part I, A clinical study of the grief of bereaved psychiatric patients. The British Journal of Medical Psychology, 38 , 1.

Pearlin, L. I. (1999). The stress process revisited. In C. S. Aneshensel & J. C. Phelan (Eds.), Handbook of sociology of mental health (pp. 395–415). New York: Kluwer Academic/Plenum.

Pearlin, L. I., Lieberman, M. A., Menaghan, E. G., & Mullan, J. T. (1981). The stress process. Journal of Health and Social Behavior, 22 , 337–356.

Pearlin, L. I., Schieman, S., Fazio, E. M., & Meersman, S. C. (2005). Stress, health, and the life course: Some conceptual perspectives. Journal of Health and Social Behavior, 46 , 205–219.

Pearlin, L. I., & Schooler, C. (1978). The structure of coping. Journal of Health and Social Behavior, 19 , 2–21.

Peterson, J., Johnson, M. A., Halvorsen, B., Apmann, L., Chang, P. C., Kershek, S., et al. (2010). What is so stressful about caring for a dying patient? A qualitative study of nurses’ experiences. International Journal of Palliative Medicine, 16 , 181–187.

Phelan, J. C., Link, B. G., & Tehranifar, P. (2010). Social conditions as fundamental causes of health inequalities: Theory, evidence, and policy implications. Journal of Health and Social Behavior, 51 , S28–S40.

Pillemer, K., & Suitor, J. J. (2008). Collective ambivalence: Considering new approaches to the complexity of intergenerational relations. Journal of Gerontology: Social Sciences, 63 , 394–396.

Price, R. H., Choi, J., & Vinokur, A. D. (2002). Links in the chain of adversity following job loss: How financial strain and loss of personal control lead to depression, impaired functioning and poor health. Journal of Occupational Health Psychology, 7 , 302–312.

Pudrovska, T. (2008). Psychological implications of motherhood and fatherhood in midlife: Evidence from sibling models. Journal of Marriage and Family, 70 , 168–181.

Pudrovska, T. (2010). Why is cancer more depressing for men than women among older white adults? Social Forces, 89 , 535–558.

Radloff, L. S. (1977). The CES-D scale: A self report depression scale for research in the general population. Applied Psychological Measurement, 1 , 385–401.

Ragin, C. C., Nagel, J., & White, P. (2004). Report of the Workshop on Scientific Foundations of Qualitative Research . Arlington, VA: National Science Foundation.

Repetti, R. L., Robles, T. F., & Reynolds, B. (2011). Allostatic processes in the family. Development and Psychopathology, 23 , 921–938.

Repetti, R. L., Taylor, S. E., & Seeman, T. E. (2002). Risky families: Family social environments and the mental and physical health of offspring. Psychological Bulletin, 128 , 330–366.

Reynolds, P., Hurley, S., Torres, M., Jackson, J., Boyd, P., & Chen, V. W. (2000). Use of coping strategies and breast cancer survival: Results from the Black/White cancer survival study. American Journal of Epidemiology, 152 , 940–949.

Robins, L. N., Wing, J., Wittchen, H. U., Helzer, J. E., Babor, T. F., Burke, J., et al. (1989). The Composite International Diagnostic Interview: An epidemiologic instrument suitable for use in conjunction with different diagnostic systems and in different cultures. Archives of General Psychiatry, 45 , 1069–1077.

Robles, T. F., & Kiecolt-Glaser, J. K. (2003). The physiology of marriage: Pathways to health. Physiology and Behavior, 79 , 409–416.

Rook, K. S. (1998). Investigating the positive and negative sides of personal relationships: Through a glass darkly? In B. H. Spitzberg & W. R. Cupach (Eds.), The dark side of close relationships (pp. 369–393). Mahwah, NJ: Lawrence Erlbaum.

Ross, C. E., & Mirowsky, J. (1989). Explaining the social patterns of depression: Control and problem solving – or support and talking. Journal of Health and Social Behavior, 30 , 206–219.

Ross, C. E., & Mirowsky, J. (2003). Social causes of psychological distress (2nd ed.). New York: Aldine.

Rotter, J. B. (1966). Generalized expectancies of internal versus external control of reinforcements. Psychological Monographs, 80 (entire issue).

Roxburgh, S. (2004). There just aren’t enough hours in the day: The mental health consequences of time pressure. Journal of Health and Social Behavior, 45 , 115–131.

Ryff, C. D. (1989). Happiness is everything, or is it? Explorations on the meaning of psychological well-being. Journal of Personality and Social Psychology, 57 , 1069–1081.

Ryff, C. D., & Singer, B. H. (2001). Emotion, social relationships, and health . New York: Oxford University Press.

Sampson, R. J., Morenoff, J. D., & Gannon-Rowley, T. (2002). Assessing “neighborhood effects”: Social processes and new directions in research. Annual Review of Sociology, 28 , 443–478.

Sandberg, J. G., Harper, J. M., Miller, R. M., Robila, M., & Davey, A. (2009). The impact of marital conflict on health and health care utilization in older couples. Journal of Health Psychology, 14 , 9–17.

Santavirta, S., Kettunen, S., & Solovieva, S. (2001). Coping in spouses of patients with acute myocardial infarction in the early phase of recovery. The Journal of Cardiovascular Nursing, 16 , 34–46.

Saxbe, D., & Repetti, R. L. (2010). For better or worse? Coregulation of couples’ cortisol levels and mood states. Journal of Personality and Social Psychology, 98 , 92–103.

Schnittker, J., & McLeod, J. D. (2005). The social psychology of health disparities. Annual Review of Sociology, 31 , 75–103.

Schoenborn, C. A., & Adams, P. F. (2010). Vital health statistics. Health behaviors of adults: United States, 2005–2007 (Series 10, Number 245) . Hyattsville, MD: National Center for Health Statistics.

Sellers, S. L., & Neighbors, H. (2008). The effects of goal-striving stress on the mental health of black Americans. Journal of Health and Social Behavior, 49 , 92–103.

Selye, H. (1956). The stress of life . New York: McGraw-Hill.

Shorter-Gooden, K. (2004). Multiple resistance strategies: How African American women cope with racism and sexism. Journal of Black Psychology, 30 , 406–425.

Simon, R. (1992). Parental role strains, salience of parental identity, and gender differences in psychological distress. Journal of Health and Social Behavior, 33 , 25–35.

Slavich, G. M., Monroe, S. M., & Gotlib, I. H. (2011). Early parental loss and depression history: Associations with recent life stress in major depressive disorder. Journal of Psychiatric Research, 45 , 1146–1152.

Slopen, N., Lewis, T. T., Gruenewald, T. L., Mujahid, M. S., Ryff, C. D., Albert, M. A., et al. (2010). Early life adversity and inflammation in African Americans and whites in the Midlife in the United States survey. Psychosomatic Medicine, 72 , 694–701.

Sorlie, P. D., Backlund, E., & Keller, J. (1995). US mortality by economic, demographic, and social characteristics: The National Longitudinal Mortality Study. American Journal of Public Health, 85 , 949–956.

Spielberger, C. D., Gorsuch, R. L., Lushene, R., Vagg, P. R., & Jacobs, G. A. (1983). Manual for the state-trait anxiety inventory . Palo Alto, CA: Consulting Psychologists Press.

Springer, K. W., Sheridan, J., Kuo, D., & Carnes, M. (2007). Long-term physical and mental health consequences of childhood physical abuse: Results from a large population-based sample of men and women. Child Abuse & Neglect, 31 , 517–530.

Steptoe, A., & Wardle, J. (2011). Positive affect measured using ecological momentary assessment and survival in older men and women. PNAS, 108 , 18244–18248.

Steptoe, A., Wardle, J., Pollard, T. M., Canaan, L., & Davies, G. J. (1996). Stress, social support, and health-related behaviour: A study of smoking, alcohol consumption and physical exercise. Journal of Psychosomatic Research, 41 , 171–180.

Taylor, S. (2010). Mechanisms linking early life stress to adult health outcomes. PNAS, 107 , 8507–8512.

Tedeschi, R. G., & Calhoun, L. G. (2004). Posttraumatic growth: Conceptual foundations and empirical evidence. Psychological Inquiry, 15 , 1–18.

Thoits, P. A. (1983). Multiple identities and psychological well-being: A reformulation and test of the social isolation hypothesis. American Sociological Review, 48 , 174–187.

Thoits, P. A. (1995). Stress, coping, and social support processes: Where are we? What next? Journal of Health and Social Behavior (Extra Issue), 53–79.

Thoits, P. A. (2010). Stress and health: Major findings and policy implications. Journal of Health and Social Behavior, 51 , S41–S53.

Turner, R. J., & Avison, W. R. (2003). Status variations in stress exposure among young adults: Implications for the interpretation of prior research. Journal of Health and Social Behavior, 44 , 488–505.

Turner, R. J., & Roszell, P. (1994). Psychosocial resources and the stress process. In W. R. Avison & I. H. Gotlib (Eds.), Stress and mental health: Contemporary issues and prospects for the future (pp. 179–210). New York: Plenum.

Turner, R. J., Wheaton, B., & Lloyd, D. A. (1995). The epidemiology of social stress. American Sociological Review, 60 , 104–125.

Twenge, J. M., & Crocker, J. (2002). Race and self-esteem: Meta-analyses comparing Whites, Blacks, Hispanics, Asians, and American Indians and comment on Gray-Little Hafdahl (2000). Psychological Bulletin, 128 , 371–408.

Umberson, D. (2003). Death of a parent: Transition to a new adult identity . Cambridge, UK: Cambridge University Press.

Umberson, D., Liu, H., & Reczek, C. (2008). Stress and health behaviors. In H. Turner & S. Schieman (Eds.), Advances in life course research: Stress processes across the life course (Vol. 13, pp. 19–44). Oxford, UK: Elsevier.

Umberson, D., Williams, K., Powers, D. A., Liu, H., & Needham, B. (2006). You make me sick: Marital quality and health over the life course. Journal of Health and Social Behavior, 47 , 1–16.

Vartanian, T. P., & Houser, L. (2010). The effects of childhood neighborhood conditions on self-reports of adult health. Journal of Health and Social Behavior, 51 , 291–306.

Wagmiller, R., Lennon, M. C., & Kuang, L. (2008). Changes in parental health and children’s economic well-being. Journal of Health and Social Behavior, 49 , 37–55.

Wallerstein, J. S., & Kelly, J. B. (1980). Surviving the breakup: How children and parents cope with divorce . New York: Basic Books.

Watson, D., Clark, L. A., & Tellegen, A. (1988). Development and validation of brief measures of positive and negative affect: PANAS scales. Journal of Psychology, 54 , 1063–1070.

Wethington, E., & Kessler, R. C. (1986). Perceived support, received support, and adjustment to stressful life events. Journal of Health and Social Behavior, 27 , 78–89.

Wheaton, B. (1990). Life transitions, role histories, and mental health. American Sociological Review, 55 , 209–223.

Wheaton, B. (1994). Sampling the stress universe. In W. R. Avison & I. H. Gotlib (Eds.), Stress and mental health: Contemporary issues and prospects for the future (pp. 77–114). New York: Plenum.

Wheaton, B. (1999). Social stress. In C. S. Aneshensel & J. Phelan (Eds.), Handbook on the sociology of mental health (pp. 277–300). New York: Plenum Press.

Whisman, M. A., & Sbarra, D. A. (2012). Marital adjustment and interlukin (IL-6). Journal of Family Psychology, 26 , 290–295.

Williams, D. R., Jackson, J., González, H. M., Neighbors, H., Nesse, R., Abelson, J. M., et al. (2007). Prevalence and distribution of major depressive disorder in African Americans, Caribbean blacks, and non-Hispanic whites. Archives of General Psychiatry, 64 , 305–315.

Williams, D. R., & Jackson, P. B. (2005). Social sources of racial disparities in health. Health Affairs, 24 , 325–334.

Williams, D. R., Yu, Y., Jackson, J. S., & Anderson, N. B. (1997). Racial differences in physical and mental health: Socioeconomic status, stress, and discrimination. Journal of Health Psychology, 2 , 335–351.

Zarit, S. H., & Zarit, J. M. (2007). Mental disorders in older adults: Fundamentals of assessment and treatment (2nd ed.). New York: Guilford Press.

Zhang, Z., & Hayward, M. (2006). Gender, the marital life course, and cardiovascular disease in late midlife. Journal of Marriage and Family, 68 , 639–657.

Zimet, G. D., Dahlem, N. W., Zimet, S. G., & Farley, G. K. (1988). The multidimensional scale of perceived social support. Journal of Personality Assessment, 52 , 30–41.

Download references

Author information

Authors and affiliations.

Department of Sociology, Rutgers University, New Brunswick, NJ, USA

Deborah Carr Ph.D.

Department of Sociology, University of Texas, Austin, TX, USA

Debra Umberson Ph.D.

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Deborah Carr Ph.D. .

Editor information

Editors and affiliations.

University of Wisconsin, Madison Dept. Sociology & Rural Sociology, MADISON, Wisconsin, USA

John DeLamater

Department of Sociology, University of Wisconsin-Madison, Madison, Wisconsin, USA

Amanda Ward

Rights and permissions

Reprints and permissions

Copyright information

© 2013 Springer Science+Business Media Dordrecht

About this chapter

Carr, D., Umberson, D. (2013). The Social Psychology of Stress, Health, and Coping. In: DeLamater, J., Ward, A. (eds) Handbook of Social Psychology. Handbooks of Sociology and Social Research. Springer, Dordrecht. https://doi.org/10.1007/978-94-007-6772-0_16

Download citation

DOI : https://doi.org/10.1007/978-94-007-6772-0_16

Published : 06 June 2013

Publisher Name : Springer, Dordrecht

Print ISBN : 978-94-007-6771-3

Online ISBN : 978-94-007-6772-0

eBook Packages : Humanities, Social Sciences and Law Social Sciences (R0)

Share this chapter

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Publish with us

Policies and ethics

  • Find a journal
  • Track your research
  • MEMBER LOGIN
  • MEDIA RELATIONS

Military Stress

research studies on stress

  • ASSESSMENTS
  • DOCUMENTARIES

research studies on stress

Stress Research

“The difficulty in science is often not so much how to make the discovery but rather to know that one has made it.”

– J.D. Bernal

Causes & Sources of Stress

Living conditions, the political climate, financial insecurity, and work issues are some stressors US adults cite as the cause of their stress. Ineffective communications increase work stress to the point of frustration that workers want to quit. These stressors, unfortunately, are not something people can just ignore. Quitting a job would result in debt and financial instability which, in turn, would be added stressors.

  • 35% of workers say their boss is a cause of their workplace stress.
  • 80% of US workers experience work stress because of ineffective company communications.
  • 39% of North American employees report their workload the main source of the work stress.
  • 49% of 18 – 24 year olds who report high levels of stress felt comparing themselves to others is a stressor.
  • 71% of US adults with private health insurance say the cost of healthcare causes them stress while 53% with public insurance say the same.
  • 54% of Americans want to stay informed about the news but following the news causes them stress.
  • 42% of US adults cite personal debt as a source of significant stress.
  • 1 in 4 American adults say discrimination is a significant source of stress.
  • Mass shootings are a significant source of stress across all races; 84% of Hispanic report this, the highest among the races.

Stress Statistics

Two years after the World Health Organization declared COVID-19 a global pandemic, inflation, money issues and the war in Ukraine have pushed U.S. stress to alarming levels, according to polls conducted for the American Psychological Association.

A late-breaking poll, fielded March 1-3 by The Harris Poll on behalf of APA, revealed striking findings, with more adults rating inflation and issues related to the invasion of Ukraine as stressors than any other issue asked about in the 15-year history of the Stress in AmericaTM poll. This comes on top of money stress at the highest recorded level since 2015, according to a broader Stress in America poll fielded last month.

Top sources of stress were the rise in prices of everyday items due to inflation (e.g., gas prices, energy bills, grocery costs, etc.) (cited by 87%), followed by supply chain issues (81%), global uncertainty (81%), Russia’s invasion of Ukraine (80%) and potential retaliation from Russia (e.g., in the form of cyberattacks or nuclear threats) (80%).

Adults also reported separation and conflict as causes for straining and/or ending of relationships. Half of adults (51%, particularly essential workers at 61%) said they have loved ones they have not been able to see in person in the past two years as a result of the COVID-19 pandemic. Strikingly, more than half of all U.S. adults (58%) reported experiencing a relationship strain or end as a result of conflicts related to the COVID-19 pandemic, including canceling events or gatherings due to COVID-19 concerns (29%); difference of opinion over some aspect of vaccines (25%); different views of the pandemic overall (25%); and difference of opinion over mask-wearing (24%).

  • 30% of Us adults eat comfort food “more than the usual” when faced with a challenging or stressful event.
  • 51% of US adults engage in prayer—a routine activity—when faced with a challenge or stressful situation.
  • Coping mechanisms of Gen Z and Millenials experiencing stress in the US 44% of Gen Z and 40% of Millenials sleep in while exercising counts for 14% and 20% respectively.
  • 49% of US adults report enduring stressful situations as a coping behavior to handle stress.
  • Less than 25% of those with depression worldwide have access to mental health treatments.

Sources: CompareCamp, American Psychological Association

Stress Management Statistics

A look at the stress management techniques employed by US adults to deal with their stress, an overwhelming majority are self-care practices. Though very helpful, it does not address the stressor at the root of the problem. Stress management programs would be beneficial not only for employees but for the company in the long run.

Stress Research from the National Library of Medicine

  • Stress and Cardiovascular Disease
  • Stress and Cancer
  • Stress and Diabetes
  • Post Traumatic Stress Disorder
  • Stress and Aging
  • Stress in Adolenscents
  • Stress and Meditation
  • Stress and Yoga
  • Workplace Stress

Cardiac Coherence and Post-traumatic Stress Disorder in Combat Veterans 

Jay P. Ginsberg, Ph.D.; Melanie E. Berry, M.S.; Donald A Powell, Ph.D.

Alternative Therapies in Health and Medicine, A Peer-Reviewed Journal, 2010;16 (4):52-60. PDF version of the complete paper: Cardiac Coherence and PTSD in Combat Veterans

The Effect of a Biofeedback-based Stress Management Tool on Physician Stress: A Randomized Controlled Clinical Trial

Jane B. Lemaire, Jean E. Wallace, Adriane M. Lewin, Jill de Grood, Jeffrey P. Schaefer

Open Medicine 2011; 5(4)E154. PDF version of the complete paper

Coherence Training In Children With Attention-Deficit Hyperactivity Disorder: Cognitive Functions and Behavioral Changes

Anthony Lloyd, Ph.D.; Davide Brett, B.Sc.; Ketith Wesnes, Ph.D.

Alternative Therapies in Health and Medicine, A Peer-Reviewed Journal, 2010; 16 (4):34-42. PDF version of the complete paper

Coherence and Health Care Cost – RCA Actuarial Study: A Cost-Effectiveness Cohort Study

Woody Bedell; Mariette Kaszkin-Bettag, Ph.D.

Alternative Therapies in Health and Medicine, A Peer-Reviewed Journal, 2010;16 (4):26-31. PDF version of the complete paper

American Psychological Association Logo

How stress affects your health

Stress can be brief, situational, and a positive force motivating performance, but if experienced over an extended period of time it can become chronic stress, which negatively impacts health and well-being.

  • Chronic Illness

How stress affects your health

Stress : We’ve all felt it. Sometimes stress can be a positive force, motivating you to perform well at your piano recital or job interview. But often—like when you’re stuck in traffic—it’s a negative force. If you experience stress over a prolonged period of time, it could become chronic—unless you take action.

A natural reaction

Have you ever found yourself with sweaty hands on a first date or felt your heart pound during a scary movie? Then you know you can feel stress in both your mind and body.

This automatic response developed in our ancient ancestors as a way to protect them from predators and other threats. Faced with danger, the body kicks into gear, flooding the body with stress hormones such as adrenaline and cortisol that elevate your heart rate, increase your blood pressure, boost your energy, and prepare you to deal with the problem.

These days, you’re not likely to face the threat of being eaten. But you probably do confront multiple challenges every day, such as meeting deadlines, paying bills, and juggling childcare that make your body react the same way. As a result, your body’s natural alarm system—the “fight or flight” response—may be stuck in the on position. And that can have serious consequences for your health.

Pressure points

Even short-lived, minor stress can have an impact. You might get a stomachache before you have to give a presentation, for example. More major acute stress, whether caused by a fight with your spouse or an event like an earthquake or terrorist attack, can have an even bigger impact.

Repeated acute stress may also contribute to inflammation in the circulatory system , particularly in the coronary arteries, and this is one pathway that is thought to tie stress to a heart attack. It also appears that how a person responds to stress can affect cholesterol levels.

Chronic stress

When stress starts interfering with your ability to live a normal life for an extended period, it becomes even more dangerous. The longer the stress lasts, the worse it is for both your mind and body. You might feel fatigued, unable to concentrate, or irritable for no good reason, for example. But chronic stress causes wear and tear on your body, too.

The long-term activation of the stress response system and the overexposure to cortisol and other stress hormones that come with it can disrupt almost all of your body's processes. This can put you at increased risk for a variety of physical and mental health problems, including anxiety, depression, digestive issues, headaches, muscle tension and pain, heart disease, heart attack, high blood pressure, stroke, sleep problems, weight gain, and memory and concentration impairment.

Chronic stress may also cause disease, either because of changes in your body or the overeating, smoking, and other bad habits people use to cope with stress. Job strain—high demands coupled with low decision-making latitude—is associated with increased risk of coronary disease , for example. Other forms of chronic stress, such as depression and low levels of social support, have also been implicated in increased cardiovascular risk.

Chronic stress also  suppresses the body's immune system , making it harder to recover from illnesses.

What you can do

Reducing your stress levels can not only make you feel better right now, but may also protect your health long-term. Several research studies have demonstrated, for example, that interventions to improve psychological health can have a beneficial impact on cardiovascular health . As a result,  researchers recommend boosting your positive affect—feelings like happiness, joy, contentment, and enthusiasm—by making time for enjoyable activities every day.

Other strategies for reducing stress include:

  • Identify what’s causing stress. Monitor your state of mind throughout the day. If you feel stressed, write down the cause, your thoughts, and your mood. Once you know what’s bothering you, develop a plan for addressing it. That might mean setting more reasonable expectations for yourself and others or asking for help with household responsibilities, job assignments, or other tasks. List all your commitments, assess your priorities, and then eliminate any tasks that are not absolutely essential.
  • Build strong relationships. Relationships can be a source of stress. Research has found that negative, hostile reactions with your spouse cause immediate changes in stress-sensitive hormones, for example. But relationships can also serve as stress buffers. Reach out to family members or close friends and let them know you’re having a tough time. They may be able to offer practical assistance and support, useful ideas, or just a fresh perspective as you begin to tackle whatever’s causing your stress.
  • Walk away when you’re angry. Before you react, take time to regroup by counting to 10. Then reconsider. Walking or other physical activities can also help you work off steam. Plus, exercise increases the production of endorphins, your body’s natural mood booster. Commit to a daily walk or other form of exercise—a small step that can make a big difference in reducing stress levels.
  • Rest your mind. To help ensure you get the recommended seven or eight hours of shut-eye, cut back on caffeine, remove distractions such as television or computers from your bedroom, and go to bed at the same time each night. Research shows that activities like yoga and relaxation exercises not only help reduce stress, but also boost immune functioning .
  • Get help. If you continue to feel overwhelmed, consult with a psychologist or other licensed mental health professional who can help you learn how to manage stress effectively. They can help you identify situations or behaviors that contribute to your chronic stress and then develop an action plan for changing them.

Recommended Reading

Waffle Can't Decide

Related reading

  • Stress effects on the body
  • Stress in America

You may also like

  • Open access
  • Published: 15 July 2024

Enhancing psychological well-being in college students: the mediating role of perceived social support and resilience in coping styles

  • Shihong Dong 1 ,
  • Huaiju Ge 1 ,
  • Wenyu Su 1 ,
  • Weimin Guan 1 ,
  • Xinquan Li 1 ,
  • Yan Liu 2 ,
  • Qing Yu 1 ,
  • Yuantao Qi 2 ,
  • Huiqing Zhang 3 &
  • Guifeng Ma 1  

BMC Psychology volume  12 , Article number:  393 ( 2024 ) Cite this article

241 Accesses

Metrics details

The prevalence of depression among college students is higher than that of the general population. Although a growing body of research suggests that depression in college students and their potential risk factors, few studies have focused on the correlation between depression and risk factors. This study aims to explore the mediating role of perceived social support and resilience in the relationship between trait coping styles and depression among college students.

A total of 1262 college students completed questionnaires including the Trait Coping Styles Questionnaire (TCSQ), the Patient Health Questionnaire-9 (PHQ-9), the Perceived Social Support Scale (PSSS), and the Resilience Scale-14 (RS-14). Common method bias tests and spearman were conducted, then regressions and bootstrap tests were used to examine the mediating effects.

In college students, there was a negative correlation between perceived control PC and depression, with a significant direct predictive effect on depression ( β = -0.067, P  < 0.01); in contrast, negative control NC showed the opposite relationship ( β  = 0.057, P  < 0.01). PC significantly positively predicted perceived social support ( β  = 0.575, P  < 0.01) and psychological resilience ( β  = 1.363, P  < 0.01); conversely, NC exerted a significant negative impact. Perceived social support could positively predict psychological resilience ( β  = 0.303, P  < 0.01), and both factors had a significant negative predictive effect on depression. Additionally, Perceived social support and resilience played a significant mediating role in the relationship between trait coping styles and depression among college students, with three mediating paths: PC/NC → perceived social support → depression among college students (-0.049/0.033), PC/NC→ resilience → depression among college students (-0.122/-0.021), and PC/NC → perceived social support → resilience → depression among college students (-0.016/0.026).

The results indicate that trait coping styles among college students not only directly predict lower depression but also indirectly influence them through perceived social support and resilience. This suggests that guiding students to confront and solve problems can alleviate their depression.

Peer Review reports

Introduction

Depression is a complex mental disorder, characterized by cognitive, affective and psychosocial symptoms [ 1 , 2 ]. It is projected that by 2030, depression will rank first globally in terms of years lived with disability [ 3 , 4 ]. Depression is also one of the most common mental health issues among contemporary college students [ 5 , 6 ]. Studies have shown that the detection rate of depression among Chinese college students ranges from 23–34% [ 7 , 8 ]. Compared to non-student populations, college students have a higher prevalence of depression, and this rate seems to be increasing [ 9 ]. This vulnerable group of college students is in a unique developmental stage, facing pressures not only from life but also from the demands of academic coursework and complex interpersonal relationships, making the factors influencing depression among college students, particularly complex [ 9 , 10 ].

Exploring the mechanisms by which influencing factors affect the occurrence of depression in college students is of significant importance for early prevention [ 11 ]. Research has demonstrated that trait coping style is one of the risk factors for depression among college students. Trait coping refers to the strategies individuals employ in challenging situations, categorized into positive coping and negative coping [ 12 , 13 ]. Positive coping focuses on taking effective action and changing stressful situations, typically associated with problem-solving behaviors and regulation of positive emotions, which can help reduce the incidence of depression [ 14 ]. Conversely, negative coping is a passive approach centered around negative evaluations and emotional expression, often involving avoiding problems and social isolation, which is more likely to lead to the development of depression [ 14 ]. Research indicates that positive coping strategies are inversely correlated with depression, serving as protective factors against depression. Conversely, negative coping strategies are positively associated with depression, acting as risk factors for its onset [ 15 ].

Perceived social support refers to an individual’s subjective emotional state of feeling supported and understood by family, friends, and other sources [ 16 , 17 ]. Prior studies have shown that perceived social support can directly impact an individual’s level of depression and also have indirect effects [ 18 ]. The data indicate that social support can significantly influence coping mechanisms, with groups having higher levels of social support tended to respond more actively and positively to stress from various sources [ 19 ]. Social support is considered an important mediating factor in determining the relationship between psychological stress and health, representing an emotional experience where individuals feel supported, respected, and understood [ 16 ]. The relationship between individuals’ coping strategies and depression may be influenced by the mediating role of perceived social support [ 20 , 21 ]. In addition to this, resilience plays a role in all three.Resilience refers to the ability to adapt to stress and adversity, enhancing an individual’s psychological well-being [ 22 ]. Both coping styles and perceived social support significantly predict resilience positively [ 23 ]. For individuals with strong resilience, possessing a high level of adaptive capacity can mitigate the negative effects of stress on individuals, thereby enhancing their mental health.

In recent years, there has been a growing body of research on the prevalence of depression among college students. However, the rates of depression vary in different environments, and there is limited research on the mechanisms through which trait coping styles, perceived social support, and resilience impact depression. Therefore, this study aims to investigate the mechanisms through which positive coping styles(PC), negative coping styles(NC), perceived social support, and resilience influence depression among college students. Additionally, it seeks to analyze the mediating roles of perceived social support and resilience in this context. The goal is to provide insights into the reasons behind depression among college students under different coping strategies, aiding in timely psychological adjustment to promote the comprehensive development of the mental and physical well-being of college students.

The following assumptions were made:

Hypothesis 1

PC has a significant negative predictive effect on depression among college students. NC has a significant positive predictive effect on depression among college students.

Hypothesis 2

Perceived social support serves as a mediator between PC/NC and depression among college students.

Hypothesis 3

Resilience mediates the relationship between PC/NC and depression among college students.

Hypothesis 4

Perceived social support and psychological resilience mediate the relationship between PC/NC and depression among college students in a serial manner.

Data and methods

This is a cross-sectional study that was conducted from January through February 2024. Using the Questionnaire Star network platform, we presented the questionnaire online, which was openly accessible to college students at a university in Shandong. The average time to complete the survey was 15 min. Participation was voluntary and students were informed about the purpose of the study. Confidentiality was assured and questionnaires were submitted anonymously. A total of 1267 enrolled college students participated in the questionnaire survey. After excluding invalid questionnaires, 1262 valid questionnaires were included, resulting in an effective rate of 99.57%.

Trait coping style questionnaire

The Trait Coping Style Questionnaire (TCSQ) [ 24 ], developed by Qianjin Jiang, was utilized to assess the trait coping styles of college students. This questionnaire reflects the participants’ approaches to coping with situations, comprising a total of 20 items. It consists of two dimensions: negative coping style and positive coping style, each with 10 items. Using a 5-point Likert scale ranging from “definitely not” to “definitely yes,” scores were assigned from 1.00 to 5.00. The Cronbach’s α coefficient for negative coping style was 0.906 and for positive coping style was 0.786 in this study.

Depression scale

The Patient Health Questionnaire-9 (PHQ-9) [ 25 ] was used to assess depressive symptoms in the past two weeks. This scale consists of 9 items rated on a 4-point Likert scale ranging from “not at all” to “nearly every day,” with scores from 0 to 3. The total score ranges from 0 to 27, with higher scores indicating more severe depressive symptoms. The Cronbach’s α coefficient for this scale in the current study was 0.884.

Perceived Social Support Scale

The Perception Social Support Scale (PSSS) was compiled by James A.Blumenthal in 1987 and later translated and modified by Qianjin Jiang to form the Chinese version of the Zimetm Perception Social Support Scale (PSSS) [ 26 , 27 ]. PSSS comprises 12 self-assessment items rated on a 7-point Likert scale. The scale includes three dimensions: family support (items 3, 4, 8, 11), friend support (items 6, 7, 9, 12), and other support (items 1, 2, 5, 10), with a total score ranging from 12 to 84. Scores of 12–36 indicate low support, 37–60 indicate moderate support, and 61–84 indicate high support. The Cronbach’s α for this scale in the current survey was 0.968.

Resilience scale

The Resilience Scale (RS-14) [ 28 ] Chinese version consists of 14 items, each rated on a 7-point Likert scale from “not at all” to “completely,” with scores ranging from 1 to 7. The total score ranges from 14 to 98, with higher scores indicating better resilience. The Cronbach’s α for this scale in the current study was 0.925.

Statistical analysis

Data were organized and analyzed using SPSS 26.0 software. Confirmatory factor analysis was first conducted on the questionnaires. Descriptive analysis was then performed on the scores of each scale. Spearman was used to examine the relationships between trait coping styles, perceived social support, resilience, and depression. Mediation analysis was carried out using the SPSS PROCESS macro 3.4.1 software model 6 developed by Hayes, specifically designed for testing complex models. Model 6 was applied for two mediating variables, followed by the bias-corrected percentile Bootstrap method with 5000 resamples to estimate the 95% confidence interval of the mediation effect. A significant mediation effect was indicated if the 95% confidence interval (CI) did not include zero. A significance level of P  < 0.05 was considered statistically significant.

Examination of common method bias

Systematic errors in indicator data results caused by the same data collection method or measurement environment can typically be assessed through the Harman single-factor test on 55 items in the dataset to examine common method bias. The results indicated that there were 7 factors with eigenvalues greater than 1, and the variance explained by the first factor was 34.84%, which was below the critical threshold of 40%. Therefore, this study may not have a significant common method bias.

Descriptive statistics and correlation analysis

The mean scores, standard deviations, and correlations of each variable are presented in Table  1 . PC ( r = -0.326, P  < 0.01), resilience ( r =-0.445, P  < 0.01), and perceived social support ( r =-0.405, P  < 0.01) were negatively correlated with depression. PC ( r  = 0.336, P  < 0.01) and resilience ( r  = 0.469, P  < 0.01) were significantly positively correlated with perceived social support. PC was significantly positively correlated with resilience( r  = 0.635, P  < 0.01). NC was significantly positively correlated with depression( r  = 0.322, P  < 0.01) and PC( r  = 0.146, P  < 0.01). NC was significantly negatively correlated with perceived social support ( r =-0.325, P  < 0.01).

Analysis of chain mediation effects

The chain mediation model was validated using SPSS PROCESS Model 6. Trait coping styles were considered as the independent variable, while depression among college students was treated as the dependent variable. Perceived social support and resilience were included as the mediating variables, culminating in the path model depicted in Figs.  1 and 2 .

The results of the regression analysis, as shown in Table  2 , indicated that PC could significantly predict perceived social support in a positive direction ( β  = 0.575, P  < 0.01). Both PC ( β  = 1.363, P  < 0.01) and perceived social support ( β  = 0.303, P  < 0.01) had significant positive predictive effects on psychological resilience. When simultaneously predicting depression using PC, perceived social support, and psychological resilience, all three exhibited significant negative predictive effects ( β = -0.067, β = -0.085, β = -0.090, P  < 0.01). NC could significantly predict perceived social support in a negative direction ( β = -0.457, P  < 0.01). When NC ( β  = 0.191, P  < 0.01) and perceived social support ( β  = 0.508, P  < 0.01) jointly predict psychological resilience, they both had significant positive predictive effects. When simultaneously predicting depression using NC, perceived social support, and psychological resilience, NC ( β  = 0.057, P  < 0.01) showed a significant positive predictive effect, while perceived social support ( β = -0.072, P  < 0.01) and psychological resilience ( β = -0.112, P  < 0.01) demonstrated significant negative predictive effects.

Further employing the Bootstrap sampling method, with 5000 repetitions, the significance of the mediating effects and chain mediation effects between trait coping styles and depression among college students was examined. The results indicated that the direct effects of PC/NC on depression were significant, with direct impact values of -0.067/0.057 (26.38%/60.00%). Perceived social support and psychological resilience mediated the relationship between PC/NC and depression, with this mediation encompassing three pathways: the separate mediating effect of perceived social support, with effect values of -0.049 and 0.033 respectively; the separate mediating effect of resilience, with effect values of -0.122 and − 0.021 respectively; and the serial mediating effect from perceived social support to resilience, with effect values of -0.016, -0.021, and 0.026. The 95% confidence intervals for all pathways did not include 0, indicating significant indirect effects. Therefore, the total indirect effects were − 0.187 (73.62%) and 0.038 (40.00%), showing that PC had a weaker direct effect on depression compared to NC, but a stronger indirect effect. This was illustrated in Table  3 .

figure 1

Chain mediation model of perceived social support and resilience between PC and depression. ** p  < 0.01

figure 2

Chain mediation model of perceived social support and resilience between NC and depression. ** p  < 0.01

Previous research on the associations and specific pathways among depressive symptoms, trait coping styles, perceived social support, and resilience in college students has been limited. Therefore, this study utilized a chain mediation model to examine how trait coping styles, perceived social support, and resilience influence depressive symptoms in college students. The results indicate that perceived social support and resilience not only act as separate mediators between PC/NC and depression but also exhibit a chain mediation effect.

Mechanisms of the impact of PC/NC on depression in college students

This study found that trait coping styles can significantly and negatively predict depressive symptoms in college students directly, consistent with previous research [ 29 ]. In recent years, amidst the backdrop of the pandemic, numerous studies have emerged domestically and internationally focusing on college students’ mental health from the perspective of crisis event coping [ 30 ]. These studies have predominantly concentrated on trait coping styles as a mediating variable in predicting the occurrence of depressive symptoms, with fewer studies examining the direct impact of trait coping styles on depressive symptoms. College students, being in a unique developmental stage, face challenges from various aspects and bear the pressures of academic coursework, interpersonal relationships, and future employment. Research indicates that trait coping styles are a key factor influencing mental health [ 31 ]. Implementing healthy coping techniques and interventions can help individuals overcome negative emotions caused by stress, which is an adaptive coping mechanism that assists college students in facing stress and enhancing problem-solving abilities, thus preventing or reducing the occurrence of depression. Conversely, adopting passive or avoidant coping strategies, leading to inadequate resolution of stress events, can increase psychological stress [ 14 ], thereby exerting a negative impact on the mental health of college students [ 32 ]. Therefore, trait coping styles play a negative predictive role in depressive symptoms among college students. PC was a positive predictor of depression and NC was a negative predictor of depression. This is consistent with previous studies [ 24 , 29 ].

Separate mediating effects of perceived social support and resilience

After introducing perceived social support and resilience as two mediating variables, the predictive effect of PC/NC on depressive symptoms in college students remained significant. The results show that PC can positively predict perceived social support, and NC is the opposite, consistent with previous research [ 33 ]. Trait coping styles are an important predictive factor in altering college students’ perceptions of social support and the occurrence of depression. Individuals who adopt negative coping styles tend to perceive relatively less external support. Some argue that social support plays a reverse predictive role in trait coping styles; the more social support college students receive and feel, the more likely they are to actively adopt positive coping strategies to alleviate stress, potentially due to variations in study subjects and time [ 34 ]. In this pathway, perceived social support can significantly and negatively predict depressive symptoms, aligning with previous research findings [ 35 ]. Perceived social support is considered a crucial mediating factor influencing mental health, referring to an individual’s ability to perceive support and understanding from family, friends, and others. College students with lower levels of perceived social support often feel neglected and undervalued, leading to negative evaluations and self-doubt, making them more susceptible to depression. PC/NC and perceived social support can interact and influence the occurrence of depressive symptoms in college students [ 16 ].

Research indicates that PC can significantly and positively predict resilience, with an indirect effect value of 48.03%.In this pathway, the mediating effect of resilience is more pronounced, consistent with previous studies [ 36 ]. There is a close connection between resilience and coping styles; college students who adopt positive coping strategies often exhibit stronger psychological resilience, being more willing to confront issues and seek help from others to solve problems. When facing pressures such as academic challenges, they approach them with a positive mindset, overcoming adversity [ 37 ]. It is believed that adopting positive coping strategies to address problems can enhance college students’ levels of psychological resilience [ 10 , 38 ]. Resilience can significantly and negatively predict depressive symptoms. depressive symptoms, College students with higher levels of resilience tend to define the severity of events less severely when stress events occur, resulting in lower psychological burdens and reduced likelihood of experiencing depressive symptoms [ 10 ]. Additionally, when facing setbacks or stress, individuals who adopt positive coping strategies actively utilize internal and external protective factors to combat current difficulties and pressures, and employ effective emotional control to mitigate the impact, thereby enhancing their levels of psychological resilience and reducing the occurrence of depression.

Chain mediation effect of perceived social support and psychological resilience

This study elucidates that PC/NC perceived social support, and psychological resilience are independent factors influencing depressive symptoms in college students, with perceived social support and psychological resilience playing a mediating role between coping styles and depressive symptoms. The share of total indirect effect values is 73.62% and 40.00%, respectively, with the third chain path accounting for 6.30% and 27.37% of the total effect ratio, respectively. This confirms the existence of this chain mediation effect, although the chain mediation effect is not as pronounced as the individual mediation effects. Positive coping styles not only directly negatively predict depressive symptoms in college students but also exert an indirect influence on depressive symptoms through perceived social support and psychological resilience. Likewise, negative coping styles not only directly positively predict depressive symptoms in college students but also have an indirect impact on depressive symptoms through perceived social support and psychological resilience, thus demonstrating the value and significance of these two mediating variables in reducing the occurrence of depressive symptoms in college students.

Initially, adopting positive coping styles and being able to perceive social support are crucial factors influencing psychological resilience in college students. There exists a relatively stable systemic relationship between students’ social support and psychological resilience, confirming that social support can enhance individuals’ levels of psychological resilience [ 16 ]. Furthermore, coping styles can affect the occurrence of depressive symptoms from both internal and external perspectives. This is because the social support perceived by college students includes not only tangible social support resources but also their subjective perception of social support, with these two factors constituting external and internal protective factors of psychological resilience [ 39 ]. Positive coping and effective adaptation can enhance college students’ perception of social support, enabling them to mobilize personal, familial, and societal protective factors better when facing various life challenges, thereby mitigating or eliminating difficulties and suppressing the onset of depressive symptoms, whereas negative coping styles yield the opposite effect. The chain mediation proposed in this study integrates the research on perceived social support, psychological resilience, and depressive symptoms in college students, facilitating a more comprehensive understanding of the internal mechanisms through which coping styles influence depressive symptoms in college students. This holds significance in advocating for a proactive attitude in college students to confront and resolve difficulties and in increasing attention to the mental health of college students.

Limitations, strengths and future research

The findings of this study hold theoretical value and practical implications, offering a reference basis for improving the mental health of college students. However, there are certain limitations to consider. Firstly, the survey in this study was conducted through self-reporting, which may introduce certain biases. Future research could explore data collection through various methods. Secondly, this study employed a cross-sectional design to investigate the impact of trait coping styles, on depression among college students and its potential mechanisms. However, this research approach does not allow for causal inferences between variables, and further validation of the study’s conclusions could be achieved through longitudinal or experimental research.

In summary, this study aims to improve the mental health of college students by examining how their coping styles, along with their perceived social support and psychological resilience, affect depressive symptoms. The research analyzes the connections between these factors and suggests that positive coping styles may help prevent depression. However, the study has its limitations and future research should use long-term experiments to better understand these relationships. Since depression in college students can be influenced by many factors, future studies should also consider additional variables and use a mix of experimental and longitudinal approaches to more clearly understand how to reduce depression in this group.

Data availability

The datasets used and analysed during the current study are available from the corresponding author upon reasonable request.

Abbreviations

Patient Health Questionnaire-9

Trait Coping Style Questionnaire

Positive coping styles

Negative coping styles

Resilience Scale

Platania GA, Savia Guerrera C, Sarti P, et al. Predictors of functional outcome in patients with major depression and bipolar disorder: a dynamic network approach to identify distinct patterns of interacting symptoms[J]. PLoS ONE. 2023;18(2):e0276822.

Article   PubMed   PubMed Central   Google Scholar  

Guerrera CS, Platania GA, Boccaccio FM, et al. The dynamic interaction between symptoms and pharmacological treatment in patients with major depressive disorder: the role of network intervention analysis[J]. BMC Psychiatry. 2023;23(1):885.

Liu Y, Chen J, Chen K, et al. The associations between academic stress and depression among college students: a moderated chain mediation model of negative affect, sleep quality, and social support[J]. Acta Psychol. 2023;239:104014.

Article   Google Scholar  

Gao L, Xie Y, Jia C, et al. Prevalence of depression among Chinese university students: a systematic review and meta-analysis[J]. Sci Rep. 2020;10(1):15897.

Naja WJ, Kansoun AH, Haddad RS. Prevalence of Depression in Medical students at the Lebanese University and exploring its correlation with Facebook Relevance: a Questionnaire Study[J]. JMIR Res Protocols. 2016;5(2):e96.

Brenneisen Mayer F, Souza Santos I, Silveira PSP, et al. Factors associated to depression and anxiety in medical students: a multicenter study[J]. BMC Med Educ. 2016;16(1):282.

Liu Y, Zhang N, Bao G, et al. Predictors of depressive symptoms in college students: a systematic review and meta-analysis of cohort studies[J]. J Affect Disord. 2019;244:196–208.

Article   PubMed   Google Scholar  

Moutinho ILD, Maddalena N, D C P, Roland RK, et al. Depression, stress and anxiety in medical students: a cross-sectional comparison between students from different semesters[J]. Volume 63. Revista da Associação Médica Brasileira; 2017. pp. 21–8. 1.

Acharya L, Jin L, Collins W. College life is stressful today – emerging stressors and depressive symptoms in college students[J]. J Am Coll Health. 2018;66(7):655–64.

Ibrahim AK, Kelly SJ, Adams CE, et al. A systematic review of studies of depression prevalence in university students[J]. J Psychiatr Res. 2013;47(3):391–400.

Zvauya R, Oyebode F, Day EJ, et al. A comparison of stress levels, coping styles and psychological morbidity between graduate-entry and traditional undergraduate medical students during the first 2 years at a UK medical school[J]. BMC Res Notes. 2017;10(1):93.

Undheim AM, Sund AM. Associations of stressful life events with coping strategies of 12–15-year-old Norwegian adolescents[J]. Eur Child Adolesc Psychiatry. 2017;26(8):993–1003.

Lau Y, Wang Y, Kwong DHK, et al. Are different coping styles Mitigating Perceived stress Associated with depressive symptoms among pregnant women? Are different coping styles Mitigating Perceived stress Associated with depressive symptoms among pregnant women?[J]. Perspect Psychiatr Care. 2016;52(2):102–12.

Ding Y, Yang Y, Yang X, et al. The Mediating Role of coping style in the relationship between Psychological Capital and Burnout among Chinese Nurses[J]. PLoS ONE. 2015;10(4):e0122128.

Gandzha IM. [Immune disorders in internal diseases and the ways for their correction][J]. Vrach Delo, 1978(7): 14–9.

Howard S, Creaven AM, Hughes BM, et al. Perceived social support predicts lower cardiovascular reactivity to stress in older adults[J]. Biol Psychol. 2017;125:70–5.

Barrera M. Distinctions between social support concepts, measures, and models[J]. Am J Community Psychol. 1986;14(4):413–45.

Xin M, Yang C, Zhang L, et al. The impact of perceived life stress and online social support on university students’ mental health during the post-COVID era in Northwestern China: gender-specific analysis[J]. BMC Public Health. 2024;24(1):467.

Sun J, Harris K, Vazire S. Is well-being associated with the quantity and quality of social interactions?[J]. J Personal Soc Psychol. 2020;119(6):1478–96.

Wang J, Chen Y, Chen H, et al. The mediating role of coping strategies between depression and social support and the moderating effect of the parent–child relationship in college students returning to school: during the period of the regular prevention and control of COVID-19[J]. Front Psychol. 2023;14:991033.

Ball S, Bax A. Self-care in Medical Education: effectiveness of health-habits interventions for First-Year Medical Students[J]. Acad Med. 2002;77(9):911–7.

Howe A, Smajdor A, Stöckl A. Towards an understanding of resilience and its relevance to medical training[J]. Med Educ. 2012;46(4):349–56.

Louise Duncan D. What the COVID-19 pandemic tells us about the need to develop resilience in the nursing workforce[J]. Nurs Manag. 2020;27(3):22–7.

Google Scholar  

Luo Y, Wang H. Correlation research on psychological health impact on nursing students against stress, coping way and social support[J]. Nurse Educ Today. 2009;29(1):5–8.

Luo Mming, Hao M, Li Xhuan, et al. Prevalence of depressive tendencies among college students and the influence of attributional styles on depressive tendencies in the post-pandemic era[J]. Front Public Health. 2024;12:1326582.

Zhang Y, Jia Y, MuLaTiHaJi M, et al. A cross-sectional mental-health survey of Chinese postgraduate students majoring in stomatology post COVID-19 restrictions[J]. Front Public Health. 2024;12:1376540.

Blumenthal JA, Burg MM, Barefoot J, et al. Social support, type A behavior, and coronary artery disease.:[J]. Psychosom Med. 1987;49(4):331–40.

Quintiliani L, Sisto A, Vicinanza F, et al. Resilience and psychological impact on Italian university students during COVID-19 pandemic. Distance learning and health[J]. Volume 27. Psychology, Health & Medicine; 2022. pp. 69–80. 1.

Shao R, He P, Ling B, et al. Prevalence of depression and anxiety and correlations between depression, anxiety, family functioning, social support and coping styles among Chinese medical students[J]. BMC Psychol. 2020;8(1):38.

Riedel B, Horen SR, Reynolds A, et al. Mental Health disorders in Nurses during the COVID-19 pandemic: implications and coping Strategies[J]. Front Public Health. 2021;9:707358.

Faisal-Cury A, Savoia MG, Menezes PR. Coping style and depressive symptomatology during pregnancy in a private setting Sample[J]. Span J Psychol. 2012;15(1):295–305.

Platania GA, Varrasi S, Guerrera CS, et al. Impact of stress during COVID-19 pandemic in Italy: a study on dispositional and behavioral dimensions for supporting evidence-based targeted Strategies[J]. Int J Environ Res Public Health. 2024;21(3):330.

Xu Y, Zheng Q, Jiang X, et al. Effects of coping on nurses’ mental health during the COVID-19 pandemic: mediating role of social support and psychological resilience[J]. Nurs Open. 2023;10(7):4619–29.

Kassam S. Understanding experiences of Social Support as Coping resources among immigrant and Refugee women with Postpartum Depression: an Integrative Literature Review[J]. Issues Ment Health Nurs. 2019;40(12):999–1011.

Xu J, Wei Y. Social Support as a moderator of the relationship between anxiety and depression: an empirical study with adult survivors of Wenchuan Earthquake[J]. PLoS ONE. 2013;8(10):e79045.

Luthar SS, Cicchetti D, Becker B. The Construct of Resilience: a critical evaluation and guidelines for future Work[J]. Child Dev. 2000;71(3):543–62.

Lee SH, Cho SJ. Cognitive Behavioral Therapy and Mindfulness-Based Cognitive Therapy for Depressive Disorders[M]//, Kim YK. Major Depressive Disorder: Vol. 1305. Singapore: Springer Singapore, 2021: 295–310.

Thompson G, McBride RB, Hosford CC, et al. Resilience among medical students: the role of coping style and social Support[J]. Teach Learn Med. 2016;28(2):174–82.

Murphy J, McGrane B, White RL, et al. Self-Esteem, meaningful experiences and the Rocky Road—contexts of physical activity that impact Mental Health in Adolescents[J]. Int J Environ Res Public Health. 2022;19(23):15846.

Download references

Acknowledgements

We would like to provide our extreme thanks and appreciation to all students who participated in our study.

This work was financially supported by the National Food Safety Risk Center Joint Research Program [grant number (LH2022GG06)] and the Weifang Medical College Teaching Reform Program (2023YBC008).

Author information

Authors and affiliations.

School of Public Health, Shandong Second Medical University, No. 7166, Baotong West Street, Weicheng District, Weifang City, 261053, China

Shihong Dong, Huaiju Ge, Wenyu Su, Weimin Guan, Xinquan Li, Qing Yu & Guifeng Ma

Shandong Cancer Research Institute (Shandong Tumor Hospital), No.440, Jiyan Road, Huaiyin District, Jinan, 250117, China

Yan Liu & Yuantao Qi

The First Affiliated Hospital of Shandong Second Medical University (Weifang People’s Hospital), No.151 Guangwen Street, Weicheng District, Weifang City, 261041, China

Huiqing Zhang

You can also search for this author in PubMed   Google Scholar

Contributions

SD and GM conceived and designed the study. HG, WS, WG, and YL undertook the data collection and analysis. SD, QY, YQ, XLand HZ drafted the manuscript. SD and GM reviewed the manuscript. The authors read and approved the final manuscript.

Corresponding authors

Correspondence to Huiqing Zhang or Guifeng Ma .

Ethics declarations

Ethics approval and consent to participate.

In accordance with the Declaration of Helsinki, the study was approved by the Ethical Committee of Shandong Second Medical University and written informed consent was required from all participants. Participation was voluntary and students were informed about the purpose of the study. Confidentiality was assured and questionnaires were submitted anonymously.

Consent for publication

Not Applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s note.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ . The Creative Commons Public Domain Dedication waiver ( http://creativecommons.org/publicdomain/zero/1.0/ ) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Cite this article.

Dong, S., Ge, H., Su, W. et al. Enhancing psychological well-being in college students: the mediating role of perceived social support and resilience in coping styles. BMC Psychol 12 , 393 (2024). https://doi.org/10.1186/s40359-024-01902-7

Download citation

Received : 04 March 2024

Accepted : 12 July 2024

Published : 15 July 2024

DOI : https://doi.org/10.1186/s40359-024-01902-7

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • College students
  • Perceived social support

BMC Psychology

ISSN: 2050-7283

research studies on stress

ORIGINAL RESEARCH article

Optimism and mental health in college students: the mediating role of sleep quality and stress.

Yun-Ju Lai

  • 1 School of Nursing, Zuckerberg College of Health Sciences, University of Massachusetts Lowell, Lowell, MA, United States
  • 2 School of Pharmacy, College of Medicine, National Taiwan University, Taipei, Taiwan
  • 3 Institute of Biomedical Sciences, Academia Sinica, Taipei, Taiwan
  • 4 Institute of Statistical Science, Academia Sinica, Taipei, Taiwan
  • 5 Department of Psychology, College of Fine Arts, Humanities & Social Sciences, University of Massachusetts Lowell, Lowell, MA, United States

Objective: College students showed a high prevalence of stress, anxiety, and depression, with medical and nursing students experiencing particularly elevated levels of mental health challenges.

Optimism significantly influences overall well-being by promoting a healthy lifestyle and cognitive responses. However, the association of optimism with sleep quality, stress, and mental health in college students remains unexplored. This study aimed to (1) explore the associations of optimism with sleep quality, stress, and mental health and (2) ascertain whether sleep quality and stress mediate the association between optimism and mental health among college students.

Methods: A cross-sectional study was conducted using online surveys with students from health science majors at a public university in the northeast United States from September to December 2022. A total of 222 students participated in the study, providing data on sociodemographics, optimism, sleep quality, stress, anxiety, and depression. Parallel and serial mediation models were utilized to examine the potential mediating roles of sleep quality and stress in the association between optimism and mental health.

Results: The study found that optimism influences anxiety and depression through both direct and indirect pathways. In line with predictions, the parallel mediation analysis revealed that the impact of optimism on anxiety (β total  = −0.598, 95% confident interval [CI]: −0.778 to −0.392) and depression (β total  = −0.724, 95% CI: −0.919 to −0.519) was mediated by stress and sleep quality. Furthermore, the serial mediation models revealed that stress and sleep quality co-mediated the relationship betweenoptimism and anxiety (indirect effect [IE] = −0.074, 95% CI: −0.135 to −0.029) or depression (IE = −0.084, 95% CI: −0.142 to −0.036) in a sequential manner.

Conclusion: Optimism was negatively correlated with poor sleep quality, stress, anxiety, and depression. Enhanced optimism was linked to high sleep quality and less stress, anxiety, and depression. These insights emphasize the potential for school-based optimism interventions to improve sleep quality, ameliorate stress-related concerns, and alleviate mental health challenges in college students.

1 Introduction

Stress, anxiety, and depression are increasingly problematic in our society, with severe consequences for both physical and mental health ( Blanco et al., 2021 ). Beyond personal health implications, mental health problems impose a substantial economic burden on society, projected to reach around 16 trillion US dollars for the global economy by 2030 ( Trautmann et al., 2016 ; Patel et al., 2018 ). College students are a particularly vulnerable population, exhibiting a higher prevalence of mental health issues compared to the general population ( Mofatteh, 2021 ). According to the National College Health Assessment from the American College Health Association, over 76% of undergraduates reported moderate or severe psychological distress ( American College Health Association, 2022 ). Furthermore, in the National Healthy Minds Study, 41% of college students disclosed their experience of major and moderate depression and 36% of them reported moderate or severe anxiety ( Healthy Minds Network, 2023 ). Poor mental health in college students may be associated with a range of problems, such as impaired academic performance, heavy alcohol consumption, substance abuse, low self-esteem, and suicide attempts ( Hysenbegasi et al., 2005 ; Skidmore et al., 2016 ; Hiçdurmaz et al., 2017 ; Liu et al., 2019 ; Sæther et al., 2019 ). Given these costly outcomes for students, universities, and society, the mental health of undergraduate students is not only a crucial public health concern but also a pressing research priority.

College students’ poor mental health may stem from various challenges, including concerns regarding academic performance, pressure to succeed, and post-graduation plans ( Beiter et al., 2015 ). Particularly, students in health science fields such as medicine and nursing often grapple with heightened levels of anxiety and depression than other non-medical peers due to their heavy workload, including theoretical responsibilities and hands-on patient care ( Mofatteh, 2021 ). In addition, the quality of sleep has been closely linked to mental health issues, with poor sleep quality exacerbating the susceptibility of college nursing students to mental illnesses, including anxiety and depression ( Zhang et al., 2018 ). Students suffering from poor sleep quality often confronted high levels of perceived stress, which in turn precipitated anxiety or depression symptoms ( Doane et al., 2015 ; Zhang et al., 2018 ), impaired psychosocial functioning ( Tavernier and Willoughby, 2014 ), and negatively impacted academic performance ( Okano et al., 2019 ). Furthermore, individuals with heightened stress were prone to develop concurrent anxiety ( Ghorbani et al., 2008 ), exhibit compromised sleep quality ( Liu et al., 2017 ), employ less healthy coping strategies ( Evans et al., 2015 ), and thus manifest depressive symptoms.

Optimism, defined as harboring positive expectations for the future ( Scheier and Carver, 1985 ), is a positive personality trait contributing to positive psychology ( Seligman and Csikszentmihalyi, 2014 ). Optimism is particularly vital during periods of uncertainty ( Carver et al., 1989 ), as demonstrated by its role as a protective factor against fear, stress, anxiety, and depression during the COVID-19 pandemic ( Vos et al., 2021 ). Heightened levels of optimism are associated with lower levels of anxiety and improved academic achievement among college students ( Singh and Jha, 2013 ), as well as better coping skills in response to stress ( Solberg Nes et al., 2009 ). Recent studies have demonstrated that optimism can promote positive emotions and higher life satisfaction, particularly during the COVID-19 pandemic ( Martinez et al., 2022 ). On the contrary, some studies suggested that lower levels of optimism and hope are aligned with decreased subjective well-being among young adults facing high levels of stress due to the pandemic ( Genç and Arslan, 2021 ).

In the present, with the majority of studies focusing on the psychological health problems in college students, the mechanisms implicated in the links among optimism, sleep quality, stress, and mental health remain unclear. Therefore, this study aimed to Blanco et al. (2021) explore the associations of optimism with sleep quality, stress, and mental health; and Trautmann et al. (2016) ascertain whether sleep and stress mediate the connection between optimism and mental health, providing valuable insights into a promising intervention strategy regarding elevating optimism levels, thereby bolstering students’ mental well-being as they confront adversity.

2 Materials and methods

2.1 study design and setting.

This study was a quantitative cross-sectional study conducted among 222 undergraduate students in health science majors at a public university in the northeast United States in the fall of 2022. We employed a non-probability purposive sampling method to administer online surveys to all freshmen, sophomore, junior, and senior health science students.

2.2 Inclusion and exclusion criteria

We included undergraduate students who were: Blanco et al. (2021) aged 18 years or older; Trautmann et al. (2016) currently enrolled full-time; and Patel et al. (2018) having access to the internet and capacity in computer typing. Students who were unable to provide informed consent were excluded.

2.3 Sample size

A minimum correlation coefficient of 0.3 was assumed for variables including optimism, sleep quality, stress, and mental health, as supported by previous studies ( Cohen, 2013 ; Zhang et al., 2018 ). With an aim to achieve 80% statistical power and a type I error rate of 0.05, a power analysis suggested a sample size of at least 67 participants for this study.

2.4 Measurements

2.4.1 revised life orientation test.

The LOT-R was employed to measure the level of optimism ( Scheier et al., 1994 ). Consisting of 10 items, the LOT-R presents with 3 items, respectively, oriented in positive and negative directions, along with 4 filler items. Respondents are requested to indicate the extent to which they agree with each item on a 5-point Likert scale that ranges from strongly disagree to strongly agree. The total score is from 0 to 24, and the higher the LOT-R score, the higher the levels of optimism. The acceptable internal consistency of the LOT-R was reported as Cronbach’s α of 0.82 ( Shifren and Anzaldi, 2018 ) in the stroke population, and stability (test–retest reliability) over a 4-month period in college students was reported as r  = 0.79 ( Scheier et al., 1994 ). A scoring range of 0 to 13 points indicates a low level of optimism, whereas a range of 14 to 18 points suggests moderate optimism, and 19 to 24 points signifies a high level of optimism ( Przybyszowski et al., 2022 ). In this study, the scale demonstrated good reliability with a Cronbach’s alpha of 0.79.

2.4.2 Pittsburgh sleep quality index

The PSQI was used to assess the sleep quality in the previous month ( Buysse et al., 1989 ). Comprising 19 questions concerning sleep habits, the PSQI involves 7 components: sleep duration, sleep latency, sleep disturbance, daytime dysfunction, use of sleeping medications, habitual sleep efficiency, and overall sleep quality. This instrument has been previously validated for college students ( Lemma et al., 2014 ). Each component is scored on a scale of 0 to 3, yielding a global sleep quality score ranging from 0 to 21. Participants with PSQI scores exceeding 5 were identified as experiencing poor sleep quality ( Buysse et al., 1989 ). The PSQI in this study demonstrated good reliability with a Cronbach’s alpha coefficient of 0.63.

2.4.3 Perceived stress scale

The PSS is a 10-item questionnaire for evaluating the perception of stress during the past month ( Cohen et al., 1983 ). Participants were asked to rate the frequency of experiencing certain feelings and thoughts using a 5-point Likert scale. With 4 positively stated items reverse scored (e.g., 0 = 4, 1 = 3, 2 = 2, 3 = 1, and 4 = 0), the total PSS scores range from 0 to 40 and the higher PSS scores indicate the higher level of stress. The scale demonstrated good internal consistency (Cronbach’s alpha) previously in undergraduate students ( Lin et al., 2020 ). The Cronbach’s alpha in this study was 0.81.

2.4.4 General anxiety disorder-7

The GAD-7 is a self-administered 7-item questionnaire, designed to assess the symptom severity of anxiety ( Spitzer et al., 2006 ). The GAD-7 is an acceptable questionnaire, with Cronbach’s alpha of 0.89 in general populations ( Lowe et al., 2008 ). Each item is scored on a 4-point Likert-type scale, ranging from 0 to 3, and summed up with a final score ranging from 0 to 21. The scores of 5, 10, and 15 are cutoff points indicating mild, moderate, and severe levels of anxiety ( Spitzer et al., 2006 ). The GAD-7 scale demonstrated excellent reliability with a Cronbach’s alpha of 0.92 for this study sample.

2.4.5 Patient health questionnaire 9

The PHQ-9 is a self-administered questionnaire, employed to test the extent of depression ( Kroenke et al., 2001 ). The internal reliability of the PHQ-9 was demonstrated, with Cronbach’s alpha of 0.81 to 0.84 ( Kroenke et al., 2016 ). Composed of 9 items, each rated on a 4-point Likert-type scale ranging from 0 to 3, the total scores on the PHQ-9 range from 0 to 27. Cutoff scores of 5, 10, 15, and 20 represent mild, moderate, moderately severe, and severe depression ( Kroenke et al., 2016 ). In this study, the scale demonstrated excellent reliability with a Cronbach’s alpha of 0.90.

2.5 Data collection

Structured questionnaires comprising students’ sociodemographics (e.g., age, biological sex, race/ethnicity, study majors, grade levels, and history of psychotherapy, psychiatric medication, or psychiatric disorder), optimism (LOT-R), sleep quality (PSQI), stress (PSS), anxiety (GAD-7), and depression (PHQ-9), were administered through Qualtrics Online Surveys. Within the online survey, all the subjects were introduced to the study’s purpose and procedures, potential risks and benefits, and assurance of privacy and confidentiality before proceeding to the survey sections. Additionally, embedded consent in the online survey required participants to agree prior to the initiation of the survey.

2.6 Data analysis

Following the assessment of statistical normality, descriptive statistics, including the number of participants, age, biological sex, race/ethnicity, study majors, and grade levels, were reported as mean ± SD or frequency (%). The Pearson correlation was applied to the relationships between sleep quality, stress, and mental health among college students. All the analyses were performed using SPSS 28.0 for Windows (SPSS Inc., Chicago, IL). Values of p  < 0.05 were considered statistically significant. Mediation analysis, for both parallel and serial mediating effects of stress and sleep quality on the relationship between optimism and mental health, was performed using the package lavaan ( Rosseel, 2012 ), version 0.6.16, implemented in the R system for statistical computing ( Team RC, 2013 ). Mediation models were adjusted for the potential covariate, biological sex, to statistically control its effects and more accurately estimate the relationship between the predictors and the outcomes in the models.

3.1 Participant characteristics

A total of 222 students participated in the study, with a mean age of 20.3 (± 2.44) years. Predominately, respondents were female (81.5%), nursing (40.5%), sophomore (33.8%), and junior (31.1%) students. Almost 20% of the participants had a psychiatric history, with 59.5 and 50.0% of them experiencing anxiety and depression, respectively ( Table 1 ). Figure 1 provides an overview of variable measurements within the respondent population. The LOT-R scores across the study population indicated a range of low optimism with a mean score of 13.12 ± 3.99. It is noteworthy that 55.3% of the students were categorized as having low optimism, while 37.7% fell into the moderate optimism category ( Table 2 ). As outlined in Table 2 and Figure 1 , participants reported experiencing various psychological states, including poor sleep quality, moderate stress, mild anxiety, and mild depression.

www.frontiersin.org

Table 1 . Descriptive statistics for respondent demographic characteristics.

www.frontiersin.org

Figure 1 . The proportion of students across various severity levels in PHQ-9, GAD-7, PSS, PSQI, and LOT-R. The X-axis indicates the percentage of students, and the Y-axis denotes the variable measurements. LOT-R, revised Life Orientation Test; PSQI, Pittsburg Sleep Quality Index; PSS, Perceived Stress Scale; GAD-7, General Anxiety Disorder-7; PHQ-9, Patient Health Questionnaire-9.

www.frontiersin.org

Table 2 . Comparison of each variable measurements by biological sex.

Upon a biological sex-based analysis, it was observed that male students exhibited notably better sleep quality, lower levels of stress, anxiety, and depression in comparison to female counterparts ( Table 2 ). However, there was no statistical difference observed in the optimism levels between male (14.08 ± 3.99) and female students (12.90 ± 3.97). In addition, no statistically significant differences were identified in the mean scores for LOT-R, PSQI, PSS, GAD-7 or PHQ-9 based on respondents’ ethnicity, study major and grade levels (Supplementary Tables 1–3).

3.2 Correlations of optimism with sleep quality, stress, and mental health among college students

As Table 3 shows, the college students’ optimism (LOT-R) demonstrated a significant negative correlation with poor sleep quality ( r  = −0.281, p  < 0.001), stress ( r  = −0.486, p  < 0.001), anxiety ( r  = −0.423, p  < 0.001), and depression ( r  = −0.476, p  < 0.001). Notably, poor sleep quality (PSQI >5) showed a positive correlation with stress ( r  = 0.498, p  < 0.001), anxiety ( r  = 0.488, p  < 0.001), and depression ( r  = 0.499, p  < 0.001). Moreover, elevated stress levels were linked to higher anxiety ( r  = 0.675, p  < 0.001) and depression ( r  = 0.675, p  < 0.001).

www.frontiersin.org

Table 3 . The associations of optimism with sleep quality, stress, anxiety, and depression among college students ( n  = 222).

3.3 The mediating role of stress and sleep quality

3.3.1 parallel mediation models.

To understand how optimism, directly and indirectly, influenced anxiety or depression, parallel mediation models were conducted. In Figure 2A , as the direct effect of optimism toward anxiety became insignificant (Direct Effect [DE] = −0.174, p  = 0.060), stress and sleep quality fully mediated the association between optimism and anxiety after the adjustment for biological sex. Specifically, stress and sleep quality exhibited significant indirect effects on this association (Indirect Effect [IE] PSS  = −0.333, p  < 0.001; IE PSQI  = −0.091, p  = 0.005). These findings indicated that students with enhanced optimism may experience lower anxiety levels through reduced stress and improved sleep quality. Besides, compared to sleep quality, lower perceived stress was the more dominant route on the impact of optimism on anxiety.

www.frontiersin.org

Figure 2 . Parallel mediation models for the mediating effects of stress and sleep quality on the relationship between optimism and mental health. Direct and indirect effects of optimism on (A) anxiety or (B) depression through stress and sleep quality are illustrated. The models are adjusted for biological sex. All effects presented are unstandardized. The direct effect and the (total effect) are depicted on the path directly from optimism to anxiety or depression. *** p  < 0.001. LOT-R, revised Life Orientation Test; PSQI, Pittsburg Sleep Quality Index; PSS, Perceived Stress Scale; GAD-7, General Anxiety Disorder-7; PHQ-9, Patient Health Questionnaire-9.

Following the adjustment for biological sex and the introduction of stress and sleep quality into the Model 2 ( Figure 2B ), the attenuated path coefficient (DE = −0.303, p  = 0.001) of optimism on depression revealed the partial mediating roles of stress and sleep quality in the optimism-depression relationship. Both stress and sleep quality demonstrated significant mediating effects on optimism with depression (IE PSS  = −0.327, p  < 0.001; IE PSQI  = −0.095, p  = 0.015), reflecting the stronger mediating role of stress. Similar to the Model 1 ( Figure 2A ), students with higher optimism levels were less susceptible to depression through reduced stress and enhanced sleep quality. Additionally, the influence of optimism on depression appeared to be primarily channeled through lower perceived stress. Overall, the total effects derived from these two models revealed statistically significant negative associations between optimism and anxiety (β total  = −0.598, 95% CI: −0.778 to −0.392), and optimism and depression (β total  = −0.724, 95% CI: −0.919 to −0.519), which implied a more pronounced impact of optimism on depression compared to anxiety.

3.3.2 Serial mediation models

To explore the mediating roles of stress and sleep quality linked in casual chains, guided by a specific directional flow, serial mediation models were employed. Since two mediators of stress and sleep quality were used and with two different outcome variables (anxiety and depression), a total of four different causal order models were produced, with only two of which were presented in Figure 3 . Each distinct casual order of the mediators was examined to compare the significant paths generated by the four models. The total indirect effects derived from serial mediation models were all found to be statistically significant. Of note, the two mediators were shown to partially mediate in the relationship between optimism and anxiety, as well as in the optimism-depression link. This observation denotes a nuanced departure from the result derived from the parallel mediation models ( Figure 2 ).

www.frontiersin.org

Figure 3 . Serial mediation models for the mediating effects of stress and sleep quality on the relationship between optimism and mental health. Four serial mediating pathways are illustrated, respectively, in the serial mediation models of (A) optimism to anxiety and (B) optimism to depression. The models are adjusted for biological sex. All effects presented are unstandardized. The direct effect and the (total effect) are depicted on the path directly from optimism to anxiety or depression. * p  < 0.05, *** p  < 0.001. LOT-R, revised Life Orientation Test; PSQI, Pittsburg Sleep Quality Index; PSS, Perceived Stress Scale; GAD-7, General Anxiety Disorder-7; PHQ-9, Patient Health Questionnaire-9.

In the analysis of serial mediation models depicting the transition from optimism to anxiety, both the indicated indirect effect path (LOT-R/PSS/PSQI/GAD-7, IE = −0.074, 95% CI: −0.135 to −0.029) presented in Figure 3A and the alternative indirect effect path (LOT-R/PSQI/PSS/GAD-7, IE = −0.066, 95% CI: −0.114 to −0.032) demonstrated statistical significance. Additionally, within the serial mediation models that trace the impact of optimism on depression, the specified indirect pathway denoted as LOT-R/PSS/PSQI/PHQ-9 ( Figure 3B , IE = −0.084, 95% CI: −0.142 to −0.036) and the alternative pathway of LOT-R/PSQI/PSS/PHQ-9 (IE = −0.064, 95% CI: −0.114 to −0.026) similarly exhibited significant mediating effect. Collectively, these findings imply that increased optimism would lead to lower perceived stress, subsequently improving sleep quality and ultimately contributing to reduced levels of anxiety and depression.

4 Discussion

This study explored the inter-relations between optimism and mental health in college students, with a specific aim to examine the mediating effects of sleep quality and stress in the relationship. Our findings revealed a significant negative correlation between optimism and poor sleep quality, stress, anxiety as well as depression. The result was consistent with the previous finding indicating that college students with higher scores on optimism reported improved sleep quality and lower levels of stress, anxiety, and depression, which might be targeted to reduce mental health problems and improve academic success ( Cheuk Yan and Wong, 2011 ; Kim et al., 2022 ). Additionally, parallel mediation analysis demonstrated that stress and sleep quality fully mediated the relationship between optimism and anxiety, while partially mediating the optimism-depression link. Regarding serial mediation analysis, a significant mediating path sequentially followed by stress and sleep quality was demonstrated both in optimism-anxiety and optimism-depression models.

With the potential underlying mechanism inferred from parallel and serial mediation analyses, this study highlighted the importance of optimism as a mechanism through which reduced levels of stress and improved sleep quality can translate into anxiety and depression. Moreover, the findings also underlined the vulnerability of health science students, as they contend with a variety of academic and clinical stressors, including long hours of study, the demanding nature of examinations, and lack of free time ( Papazisis et al., 2008 ; Sakai et al., 2022 ). Therefore, school-based interventions may hold promise in ameliorating students’ stress, improving their sleep quality, and further reducing the levels of anxiety, and depression.

Aside from examining the association between optimism and mental health among college students, this study also serves to confirm the degree of optimism, sleep quality, stress, anxiety, and depression experienced by college students in the post-COVID era. The emergence of the coronavirus disease 2019 (COVID-19) has brought forth not only physical health problems but also psychological issues ( Tran et al., 2020 ; Xiong et al., 2020 ). College students, in particular, are adversely impacted owing to the uncertainty surrounding academic achievement, future career prospects, and social lives ( Aristovnik et al., 2021 ). The disruptions caused by school closures, cancelation of social events, remote online courses, and exam postponements during the COVID-19 pandemic heightened their emotional distress ( Cao et al., 2020 ; Lei et al., 2020 ; Xiong et al., 2020 ). However, the psychological challenges faced by students did not end with the remission of the pandemic and the easing of social restrictions. In this study, our participants reported low optimism, poor sleep quality, moderate stress, mild anxiety, and depression, reflecting the persisting psychological impact of the pandemic. Consistent with our findings, other study has also highlighted that students perceived intensified levels of stress and anxiety, as well as moderate depression after returning to campus ( Al-Rawi et al., 2022 ). Factors contributing to their psychological struggles include fear of infection and enduring social, family, and economic changes resulting from the COVID-19 pandemic ( Wang et al., 2022 ).

There are several limitations inherent in this study. First, the study sample was exclusively drawn from Health Sciences students at a single public university in the United States, thereby limiting the generalizability of the findings to broader populations or countries. Second, the biological sex composition of the participants was predominately female, and recent studies have shown biological sex-based differences in student mental health consequences of the COVID-19 pandemic ( Ausín et al., 2021 ; Prowse et al., 2021 ). Hence, this biological sex imbalance may limit the generalizability of the results to the male population. Third, the study lacked longitudinal follow-up and, therefore, did not demonstrate causal relationships, although the outcomes do imply that researchers, clinicians, and schools should take into account these variable interactions between optimism and mental health among college students.

5 Conclusion

This study documented the direct and indirect effects of stress and sleep quality and its sequential mediating pathway in the connection between optimism and mental health within health science college students. Findings from the study underscore the significance of fostering academic optimism to alleviate stress and improve sleep quality, ultimately expecting to ease the mental health burdens experienced by college students. Consequently, the development of diverse academic programs focused on enhancing the optimism of college students becomes imperative.

Data availability statement

The raw data supporting the conclusions of this article will be made available by the authors, without undue reservation.

Ethics statement

The study involving human participants was reviewed and approved by the Institutional Review Board (IRB) at the University of Massachusetts Lowell (IRB# 22-072-LAI-EXM). The studies were conducted in accordance with the local legislation and institutional requirements. The participants provided their written informed consent to participate in this study.

Author contributions

Y-JL: Conceptualization, Data curation, Formal analysis, Funding acquisition, Investigation, Methodology, Project administration, Resources, Supervision, Validation, Visualization, Writing – original draft, Writing – review & editing. E-YT: Resources, Visualization, Writing – review & editing. PJ: Formal analysis, Writing – original draft, Writing – review & editing. Y-SW: Formal analysis, Methodology, Writing – review & editing. Y-HC: Formal analysis, Methodology, Writing – review & editing. SO’L: Data curation, Writing – review & editing. SM: Data curation, Writing – review & editing. YZ: Conceptualization, Validation, Writing – review & editing. JS: Validation, Writing – review & editing. YW: Methodology, Validation, Writing – review & editing.

The author(s) declare that financial support was received for the research, authorship, and/or publication of this article. This study was funded by the Donna Manning Pilot Research Program and the University of Massachusetts Lowell (Faculty start-up D50210000000022 from Y-JL).

Acknowledgments

The authors express gratitude to all the participants in this study.

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of int.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Supplementary material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fpsyg.2024.1403146/full#supplementary-material .

Al-Rawi, S. S., Jumah, H. A., Ibrahim, A. H., Hamdy, H. A., Hama, H. A., Abdulrahman, M. D., et al. (2022). Depression, anxiety and stress level among university students of class reentry post Covid-19 pandemic. International journal of social sciences & educational. Aust. Stud. 9:197. doi: 10.23918/ijsses.v9i2p197

Crossref Full Text | Google Scholar

American College Health Association (2022). American college health association-National College Health Assessment III . Silver Spring, Maryland: Undergraduate Student Reference Group Data Report Spring.

Google Scholar

Aristovnik, A., Keržič, D., Ravšelj, D., Tomaževič, N., and Umek, L. (2021). Impacts of the Covid-19 pandemic on life of higher education students: global survey dataset from the first wave. Data Brief 39:107659. doi: 10.1016/j.dib.2021.107659

PubMed Abstract | Crossref Full Text | Google Scholar

Ausín, B., González-Sanguino, C., Castellanos, M. Á., and Muñoz, M. (2021). Gender-related differences in the psychological impact of confinement as a consequence of COVID-19 in Spain. J. Gend. Stud. 30, 29–38. doi: 10.1080/09589236.2020.1799768

Beiter, R., Nash, R., McCrady, M., Rhoades, D., Linscomb, M., Clarahan, M., et al. (2015). The prevalence and correlates of depression, anxiety, and stress in a sample of college students. J. Affect. Disord. 173, 90–96. doi: 10.1016/j.jad.2014.10.054

Blanco, V., Salmerón, M., Otero, P., and Vázquez, F. L. (2021). Symptoms of depression, anxiety, and stress and prevalence of major depression and its predictors in female university students. Int. J. Environ. Res. Public Health 18:5845. doi: 10.3390/ijerph18115845

Buysse, D. J., Reynolds, C. F., Monk, T. H., Berman, S. R., and Kupfer, D. J. (1989). The Pittsburgh sleep quality index: a new instrument for psychiatric practice and research. Psychiatry Res. 28, 193–213. doi: 10.1016/0165-1781(89)90047-4

Cao, W., Fang, Z., Hou, G., Han, M., Xu, X., Dong, J., et al. (2020). The psychological impact of the COVID-19 epidemic on college students in China. Psychiatry Res. 287:112934. doi: 10.1016/j.psychres.2020.112934

Carver, C. S., Scheier, M. F., and Weintraub, J. K. (1989). Assessing coping strategies: a theoretically based approach. J. Pers. Soc. Psychol. 56, 267–283. doi: 10.1037/0022-3514.56.2.267

Cheuk Yan, S., and Wong, W. S. (2011). The effect of optimism on depression: the mediating and moderating role of insomnia. J. Health Psychol. 16, 1251–1258. doi: 10.1177/1359105311407366

Cohen, S., Kamarck, T., and Mermelstein, R. (1983). A global measure of perceived stress. J. Health Soc. Behav. 24, 385–396. doi: 10.2307/2136404

Cohen, (2013). Statistical power analysis for the behavioral sciences . New York: Routledge.

Doane, L. D., Gress-Smith, J. L., and Breitenstein, R. S. (2015). Multi-method assessments of sleep over the transition to college and the associations with depression and anxiety symptoms. J. Youth Adolesc. 44, 389–404. doi: 10.1007/s10964-014-0150-7

Evans, L. D., Kouros, C., Frankel, S. A., McCauley, E., Diamond, G. S., Schloredt, K. A., et al. (2015). Longitudinal relations between stress and depressive symptoms in youth: coping as a mediator. J. Abnorm. Child Psychol. 43, 355–368. doi: 10.1007/s10802-014-9906-5

Genç, E., and Arslan, G. (2021). Optimism and dispositional hope to promote college students’ subjective well-being in the context of the COVID-19 pandemic. J. Positive School Psychol. 5, 87–96. doi: 10.47602/jpsp.v5i2.255

Ghorbani, N., Krauss, S. W., Watson, P. J., and Lebreton, D. (2008). Relationship of perceived stress with depression: complete mediation by perceived control and anxiety in Iran and the United States. J. international de psychologie. 43, 958–968. doi: 10.1080/00207590701295264

Healthy Minds Network (2023). Healthy minds study among colleges and universities, 2022–2023 . Los Angeles: Healthy minds network, University of Michigan, University of California, Boston University, and Wayne State University.

Hiçdurmaz, D., Inci, F., and Karahan, S. (2017). Predictors of mental health symptoms, automatic thoughts, and self-esteem among university students. Psychol. Rep. 120, 650–669. doi: 10.1177/0033294117707945

Hysenbegasi, A., Hass, S. L., and Rowland, C. R. (2005). The impact of depression on the academic productivity of university students. J. Ment. Health Policy Econ. 8, 145–151

PubMed Abstract | Google Scholar

Kim, M.-J., Shin, G.-Y., and Choi, Y.-S. (2022). The effects of depression and optimism on academic stress in Korean university students in COVID-19 situation. J. ReAttach Therapy and Develop. Diversities. 5, 352–363.

Kroenke, K., Spitzer, R. L., and Williams, J. B. (2001). The PHQ-9: validity of a brief depression severity measure. J. Gen. Intern. Med. 16, 606–613. doi: 10.1046/j.1525-1497.2001.016009606.x

Kroenke, K., Wu, J., Yu, Z., Bair, M. J., Kean, J., Stump, T., et al. (2016). Patient health questionnaire anxiety and depression scale: initial validation in three clinical trials. Psychosom. Med. 78, 716–727. doi: 10.1097/PSY.0000000000000322

Lei, L., Huang, X., Zhang, S., Yang, J., Yang, L., and Xu, M. (2020). Comparison of prevalence and associated factors of anxiety and depression among people affected by versus people unaffected by quarantine during the COVID-19 epidemic in southwestern China. Med. Sci. Monitor: Int. Med J. Experimental Clin. Res. 26:e924609. doi: 10.12659/MSM.924609

Lemma, S., Berhane, Y., Worku, A., Gelaye, B., and Williams, M. A. (2014). Good quality sleep is associated with better academic performance among university students in Ethiopia. Sleep Breath. 18, 257–263. doi: 10.1007/s11325-013-0874-8

Lin, X. J., Zhang, C. Y., Yang, S., Hsu, M. L., Cheng, H., Chen, J., et al. (2020). Stress and its association with academic performance among dental undergraduate students in Fujian, China: a cross-sectional online questionnaire survey. BMC Med. Educ. 20:181. doi: 10.1186/s12909-020-02095-4

Liu, Y., Li, T., Guo, L., Zhang, R., Feng, X., and Liu, K. (2017). The mediating role of sleep quality on the relationship between perceived stress and depression among the elderly in urban communities: a cross-sectional study. Public Health 149, 21–27. doi: 10.1016/j.puhe.2017.04.006

Liu, C. H., Stevens, C., Wong, S. H., Yasui, M., and Chen, J. A. (2019). The prevalence and predictors of mental health diagnoses and suicide among US college students: implications for addressing disparities in service use. Depress. Anxiety 36, 8–17. doi: 10.1002/da.22830

Lowe, B., Decker, O., Muller, S., Brahler, E., Schellberg, D., Herzog, W., et al. (2008). Validation and standardization of the generalized anxiety disorder screener (GAD-7) in the general population. Med. Care 46, 266–274. doi: 10.1097/MLR.0b013e318160d093

Martinez, L., Valenzuela, L. S., and Soto, V. E. (2022). Well-being amongst college students during COVID-19 pandemic: evidence from a developing country. Int. J. Environ. Res. Public Health 19:16745. doi: 10.3390/ijerph192416745

Mofatteh, M. (2021). Risk factors associated with stress, anxiety, and depression among university undergraduate students. AIMS Public Health. 8, 36–65. doi: 10.3934/publichealth.2021004

Okano, K., Kaczmarzyk, J. R., Dave, N., Gabrieli, J. D. E., and Grossman, J. C. (2019). Sleep quality, duration, and consistency are associated with better academic performance in college students. Sci. Learn. 4:16. doi: 10.1038/s41539-019-0055-z

Papazisis, G., Vlasiadis, I., Papanikolaou, N., Tsiga, E., and Sapountzi-Krepia, D. (2008). Depression and anxiety among nursing students in Greece. Ann. General Psychiatry 7:S209. doi: 10.1186/1744-859X-7-S1-S209

Patel, V., Saxena, S., Lund, C., Thornicroft, G., Baingana, F., Bolton, P., et al. (2018). The lancet commission on global mental health and sustainable development. Lancet 392, 1553–1598. doi: 10.1016/S0140-6736(18)31612-X

Prowse, R., Sherratt, F., Abizaid, A., Gabrys, R. L., Hellemans, K. G. C., Patterson, Z. R., et al. (2021). Coping with the COVID-19 pandemic: examining gender differences in stress and mental health among university students. Front. Psychol. 12:650759. doi: 10.3389/fpsyt.2021.650759

Przybyszowski, M., Pilinski, R., Sliwka, A., Polczyk, R., Nowobilski, R., Sladek, K., et al. (2022). The impact of clinical and psychological factors on asthma control: the experience of a single asthma center in Poland. J. Asthma 59, 407–417. doi: 10.1080/02770903.2020.1841791

Rosseel, Y. (2012). Lavaan: an R package for structural equation modeling. J. Stat. Softw. 48, 1–36. doi: 10.18637/jss.v048.i02

Sæther, S. M. M., Knapstad, M., Askeland, K. G., and Skogen, J. C. (2019). Alcohol consumption, life satisfaction and mental health among Norwegian college and university students. Addict. Behav. Rep. 10:100216. doi: 10.1016/j.abrep.2019.100216

Sakai, M., Nakanishi, M., Yu, Z., Takagi, G., Toshi, K., Wakashima, K., et al. (2022). Depression and anxiety among nursing students during the COVID-19 pandemic in Tohoku region, Japan: a cross-sectional survey. Japan J. Nurs. Sci. JJNS. 19:e12483. doi: 10.1111/jjns.12483

Seligman, M. E. P., and Csikszentmihalyi, M. (2014). Positive psychology: An introduction. Flow and the foundations of positive psychology: The collected works of Mihaly Csikszentmihalyi , Dordrecht, Netherlands: Springer. 279–298.

Scheier, M. F., and Carver, C. S. (1985). Optimism, coping, and health: assessment and implications of generalized outcome expectancies. Health Psychol. 4, 219–247. doi: 10.1037/0278-6133.4.3.219

Scheier, M. F., Carver, C. S., and Bridges, M. W. (1994). Distinguishing optimism from neuroticism (and trait anxiety, self-mastery, and self-esteem): a reevaluation of the life orientation test. J. Pers. Soc. Psychol. 67, 1063–1078. doi: 10.1037/0022-3514.67.6.1063

Shifren, K., and Anzaldi, K. (2018). Optimism, well-being, depressive symptoms, and perceived physical health: a study among stroke survivors. Psychol. Health Med. 23, 46–57. doi: 10.1080/13548506.2017.1325505

Singh, I., and Jha, A. (2013). Anxiety, optimism and academic achievement among students of private medical and engineering colleges: a comparative study. Journal of Educational and Developmental Psychology. 3: 222.

Skidmore, C. R., Kaufman, E. A., and Crowell, S. E. (2016). Substance use among college students. Child Adoles. Psychiatric Clin. 25, 735–753. doi: 10.1016/j.chc.2016.06.004

Solberg Nes, L., Evans, D. R., and Segerstrom, S. C. (2009). Optimism and college retention: mediation by motivation, performance, and adjustment 1. J. Appl. Soc. Psychol. 39, 1887–912.

Spitzer, R. L., Kroenke, K., Williams, J. B., and Lowe, B. (2006). A brief measure for assessing generalized anxiety disorder: the GAD-7. Arch. Intern. Med. 166, 1092–1097. doi: 10.1001/archinte.166.10.1092

Tavernier, R., and Willoughby, T. (2014). Bidirectional associations between sleep (quality and duration) and psychosocial functioning across the university years. Dev. Psychol. 50, 674–682. doi: 10.1037/a0034258

Team RC. R . A language and environment for statistical computing. R Foundation for Statistical Computing . Vienna, Austria (2021).

Tran, B. X., Ha, G. H., Nguyen, L. H., Vu, G. T., Hoang, M. T., Le, H. T., et al. (2020). Studies of novel coronavirus disease 19 (COVID-19) pandemic: a global analysis of literature. Int. J. Environ. Res. Public Health 17:4095. doi: 10.3390/ijerph17114095

Trautmann, S., Rehm, J., and Wittchen, H. U. (2016). The economic costs of mental disorders: do our societies react appropriately to the burden of mental disorders? EMBO Rep. 17, 1245–1249. doi: 10.15252/embr.201642951

Vos, L. M., Habibović, M., Nyklíček, I., Smeets, T., and Mertens, G. (2021). Optimism, mindfulness, and resilience as potential protective factors for the mental health consequences of fear of the coronavirus. Psychiatry Res. 300:113927. doi: 10.1016/j.psychres.2021.113927

Wang, X., Zhang, N., Pu, C., Li, Y., Chen, H., and Li, M. (2022). Anxiety, depression, and PTSD among college students in the post-COVID-19 era: a cross-sectional study. Brain Sci. 12:1553. doi: 10.3390/brainsci12111553

Xiong, J., Lipsitz, O., Nasri, F., Lui, L. M., Gill, H., Phan, L., et al. (2020). Impact of COVID-19 pandemic on mental health in the general population: a systematic review. J. Affect. Disord. 277, 55–64. doi: 10.1016/j.jad.2020.08.001

Zhang, Y., Peters, A., and Chen, G. (2018). Perceived stress mediates the associations between sleep quality and symptoms of anxiety and depression among college nursing students. Int. J. Nurs. Educ. Scholarsh. 15:20170020. doi: 10.1515/ijnes-2017-0020

Keywords: optimism, sleep quality, stress, anxiety, depression, mediation analysis

Citation: Lai Y-J, Tsai E-Y, Jarustanaput P, Wu Y-S, Chen Y-H, O’Leary SE, Manachevakul S, Zhang Y, Shen J and Wang Y (2024) Optimism and mental health in college students: the mediating role of sleep quality and stress. Front. Psychol . 15:1403146. doi: 10.3389/fpsyg.2024.1403146

Received: 18 March 2024; Accepted: 25 June 2024; Published: 16 July 2024.

Reviewed by:

Copyright © 2024 Lai, Tsai, Jarustanaput, Wu, Chen, O’Leary, Manachevakul, Zhang, Shen and Wang. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY) . The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Yun-Ju Lai, [email protected]

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.

NIMH Logo

Transforming the understanding and treatment of mental illnesses.

Información en español

Celebrating 75 Years! Learn More >>

  • Health Topics
  • Brochures and Fact Sheets
  • Help for Mental Illnesses
  • Clinical Trials

Post-Traumatic Stress Disorder (PTSD)

Post-traumatic stress disorder (PTSD) can develop after exposure to a potentially traumatic event that is beyond a typical stressor. Events that may lead to PTSD include, but are not limited to, violent personal assaults, natural or human-caused disasters, accidents, combat, and other forms of violence. Exposure to events like these is common. About one half of all U.S. adults will experience at least one traumatic event in their lives, but most do not develop PTSD. People who experience PTSD may have persistent, frightening thoughts and memories of the event(s), experience sleep problems, feel detached or numb, or may be easily startled. In severe forms, PTSD can significantly impair a person's ability to function at work, at home, and socially.

Additional information about PTSD can be found on the NIMH Health Topics page on Post-Traumatic Stress Disorder .

Prevalence of Post-Traumatic Stress Disorder Among Adults

  • An estimated 3.6% of U.S. adults had PTSD in the past year.
  • Past year prevalence of PTSD among adults was higher for females (5.2%) than for males (1.8%).
  • The lifetime prevalence of PTSD was 6.8%. 2
Past Year Prevalence of Post-Traumatic Stress Disorder Among Adults (2001-2003)
Demographic Percent
Overall 3.6
Sex Female 5.2
Male 1.8
Age 18-29 4.0
30-44 3.5
45-59 5.3
60+ 1.0

Post-Traumatic Stress Disorder with Impairment Among Adults

  • Of adults with PTSD in the past year, degree of impairment ranged from mild to serious, as shown in Figure 2. Impairment was determined by scores on the Sheehan Disability Scale. 3
  • Impairment was distributed evenly among adults with PTSD. An estimated 36.6% had serious impairment, 33.1% had moderate impairment, and 30.2% had mild impairment
Past Year Severity of Post-Traumatic Stress Disorder Among U.S. Adults (2001-2003)
Severity Percent
Mild 30.2
Moderate 33.1
Serious 36.6
Total 100.0

Prevalence of Post-Traumatic Stress Disorder Among Adolescents

  • An estimated 5.0% of adolescents had PTSD, and an estimated 1.5% had severe impairment. Diagnostic and Statistical Manual of Mental Disorders, Fourth Edition (DSM-IV) criteria were used to determine impairment.
  • The prevalence of PTSD among adolescents was higher for females (8.0%) than for males (2.3%).
Lifetime Prevalence of Post-Traumatic Stress Disorder Among Adolescents (2001-2004)
Demographic Percent
Overall 5.0
With Severe Impairment 1.5
Sex Female 8.0
Male 2.3
Age 13-14 3.7
15-16 5.1
17-18 7.0

Data Sources

  • Harvard Medical School, 2007. National Comorbidity Survey (NCS). (2017, August 21). Retrieved from https://www.hcp.med.harvard.edu/ncs/index.php   . Data Table 2: 12-month prevalence DSM-IV/WMH-CIDI disorders by sex and cohort   .
  • Harvard Medical School, 2007. National Comorbidity Survey (NCS). (2017, August 21). Retrieved from https://www.hcp.med.harvard.edu/ncs/index.php   . Data Table 1: Lifetime prevalence DSM-IV/WMH-CIDI disorders by sex and cohort   .
  • Kessler RC, Chiu WT, Demler O, Merikangas KR, Walters EE. Prevalence, severity, and comorbidity of 12-month DSM-IV disorders in the National Comorbidity Survey Replication. Arch Gen Psychiatry. 2005 Jun;62(6):617-27. PMID: 15939839 
  • Merikangas KR, He JP, Burstein M, Swanson SA, Avenevoli S, Cui L, Benjet C, Georgiades K, Swendsen J. Lifetime prevalence of mental disorders in U.S. adolescents: results from the National Comorbidity Survey Replication--Adolescent Supplement (NCS-A). J Am Acad Child Adolesc Psychiatry. 2010 Oct;49(10):980-9. PMID: 20855043 

Statistical Methods and Measurement Caveats

National comorbidity survey replication (ncs-r).

Diagnostic Assessment and Population:

  • The NCS-R is a nationally representative, face-to-face, household survey conducted between February 2001 and April 2003 with a response rate of 70.9%. DSM-IV mental disorders were assessed using a modified version of the fully structured World Health Organization Composite International Diagnostic Interview (WMH-CIDI), a fully structured lay-administered diagnostic interview that generates both International Classification of Diseases, 10 th Revision, and DSM-IV diagnoses. The DSM-IV criteria were used here. The Sheehan Disability Scale (SDS) assessed disability in work role performance, household maintenance, social life, and intimate relationships on a 0–10 scale. Participants for the main interview totaled 9,282 English-speaking, non-institutionalized, civilian respondents. Post-traumatic stress disorder (PTSD) was assessed in a subsample of 5,692 adults. The NCS-R was led by Harvard University.
  • Unlike the DSM-IV criteria used in the NCS-R and NCS-A, the current DSM-5 no longer places PTSD in the anxiety disorder category. It is listed in a new DSM-5 category, Trauma- and Stressor-Related Disorders.

Survey Non-response:

  • In 2001-2002, non-response was 29.1% of primary respondents and 19.6% of secondary respondents.
  • Reasons for non-response to interviewing include: refusal to participate (7.3% of primary, 6.3% of secondary); respondent was reluctant- too busy but did not refuse (17.7% of primary, 11.6% of secondary); circumstantial, such as intellectual developmental disability or overseas work assignment (2.0% of primary, 1.7% of secondary); and household units that were never contacted (2.0).
  • For more information, see PMID: 15297905  .

National Comorbidity Survey Adolescent Supplement (NCS-A)

  • The NCS-A was carried out under a cooperative agreement sponsored by NIMH to meet a request from Congress to provide national data on the prevalence and correlates of mental disorders among U.S. youth. The NCS-A was a nationally representative, face-to-face survey of 10,123 adolescents aged 13 to 18 years in the continental United States. The survey was based on a dual-frame design that included 904 adolescent residents of the households that participated in the adult U.S. National Comorbidity Survey Replication and 9,244 adolescent students selected from a nationally representative sample of 320 schools. The survey was fielded between February 2001 and January 2004. DSM-IV mental disorders were assessed using a modified version of the fully structured World Health Organization Composite International Diagnostic Interview.
  • The overall adolescent non-response rate was 24.4%. This is made up of non-response rates of 14.1% in the household sample, 18.2% in the un-blinded school sample, and 77.7% in the blinded school sample. Non-response was largely due to refusal (21.3%), which in the household and un-blinded school samples came largely from parents rather than adolescents (72.3% and 81.0%, respectively). The refusals in the blinded school sample, in comparison, came almost entirely (98.1%) from parents failing to return the signed consent postcard.
  • For more information, see PMID: 19507169  .
  • Contact DTN

DTN Progressive Farmer

  • World & Policy
  • Business & Inputs

DTN Oil Update

This information is not available at this time. We will update as soon as possible.

Farm Bill, Extension Discussed at RNC

Gop congressional aggies talk about farm bill, trade at rnc, top 5 things to watch.

(DTN/Progressive Farmer graphic)

What Condition Our Condition is In

Crop and field reports, and the repercussions of political events are all making us drop in to see what condition our condition is in. Cue the...

Ag Seeks SAF Feedstock Mandate on 45Z

Ag groups want irs to mandate domestic feedstock use to qualify for saf credit, michigan farmer wants fraud dismissed, michigan farmer motions to dismiss wire fraud, tells court evidence lacking, timing important for proper hay baling, moisture causes challenges for baling hay properly, view from the cab.

This Idaho flax field may look like a painting, but Dan Lakey is trying not to look too hard at Mother Nature's brush strokes this season. Weather has taken a toll on many of his crops. (DTN photo courtesy of Dan Lakey)

Farmers Discuss Mending Fences, Roads, Machinery and Life

Weather continues to create interesting scenarios for DTN's View From the Cab farmers in Idaho and Kentucky. This week updates on some...

Study: Link Between Ag Stress, Alcohol

Study finds 27% of farmers surveyed report binge drinking to alleviate stress, farm crops face hot, dry and uncertain summer, rebuilding rural resilience in recovery, rural leaders highlight disaster recovery challenges.

  • Special Sections
  • Past Issues
  • LOG IN or JOIN OUR COMMUNITY

Todd Neeley

LINCOLN, Neb. (DTN) -- About one in four farmers in the United States turn to binge drinking in response to high stress levels, according to a new study conducted by researchers at the University of Georgia.

What's more, the study that included a survey of 1,045 farmers found them turning to alcohol because of stigma among the farming community and actual barriers to accessing mental health care in rural areas.

"Beyond cultural factors that act as a barrier to help-seeking, like resilience and the stigmatization of mental health in the farming community, farmers are also isolated from both social support and healthcare resources," the study published in the Journal of Agromedicine said.

"These factors may reinforce the use of alcohol to manage the high levels of stress associated with farming. The risk associated with farming, lack of access to healthcare resources, and a scarcity of healthcare providers who understand the farming population, are compounded by stigma associated with seeking treatment for mental health and substance use disorders creating a vicious cycle that promotes unhealthy coping strategies in farming populations."

The survey was distributed to the ag community in a number of ways including at the 2023 Georgia and American Farm Bureau Federation conferences between Nov. 1, 2022, and Feb. 1, 2023.

A whopping 96% of survey respondents reported consuming alcohol. That includes about 27% of farmers reporting drinking alcohol two to three times or four or more times a week. About 35% of farmers surveyed reported consuming alcohol roughly two to four times per month. About 70% of respondents were male while 28% of respondents were female, with an average age of about 33.

"When participants consumed alcohol, 34% indicated that they consumed three or four drinks in one sitting, 22.5% reported drinking five or six drinks in one sitting, with 18.2% reported having seven or more drinks in one sitting over the last three months," the paper said.

"Binge drinking behavior was reported on a weekly or daily basis by 23.4% of respondents, with 19.6% reporting binge drinking on a weekly basis and 3.8% reporting binge drinking on a daily basis."

FARMER STRESS LEVELS

The study authors also surveyed producers about their perceived stress and mental health stigma. The farmers surveyed reported overall high stress levels, as well as concerns about social stigma and financial costs connected with seeking mental health services.

"Participants most strongly endorsed concerns about others taking them less seriously, treating them differently and talking about them behind their back, as well as personal feelings of shame as markers of stigma associated with help-seeking for mental health support," the study authors said in the published paper.

"Participants endorsed a fair amount of formal healthcare challenges. Participants most strongly endorsed concerns about paying out of pocket for specialty care, lack of insurance, and lack of awareness of available resources for substance use disorders as primary challenges associated with accessing healthcare."

The farmers surveyed reported being either second- or third-generation producers and 80% of them are married. Most of them, about 48%, reported being farm owners or managers. About 78% of the farmers surveyed produce beef cattle, wheat and corn.

STUDY LIMITATIONS

The study authors outline a number of limitations with their work including that survey respondents were much younger than the average age of farmers across the industry.

For instance, the younger demographic surveyed could be attributed to the use of digital recruitment strategies as well as the content of the survey itself, the authors said.

"Participants in prior research conducted in the farming community indicated that older farmers would be unlikely to engage in discussions of stressors and mental health, and this generational distinction may have contributed to a younger respondent pool," according to the paper.

"Within the farming population, it has been observed that younger farmers are more likely to report not only higher stress but higher rates of alcohol consumption as well which might have skewed the results of this study. In addition, a majority of responses were from farm owners and managers with limited responses from farm workers and spouses.

The study authors said farm workers are more likely to be foreign-born, "experience poverty," lack access to health insurance and have lower education levels which could influence behavior and health outcomes.

Christina Proctor, lead author of the study and a clinical associate professor at UGA's College of Public Health, said in a news release that alcohol is the "most acceptable way" for farmers to deal with stress rather than seeking out help.

Statistics from the Centers for Disease Control and Prevention show suicide rates in rural areas are higher than anywhere else in the country.

"Knowing the stigma that exists within rural farming populations about seeking care and then looking at death by suicide numbers, it really is a public health issue because there are drastic, traumatic outcomes associated with not being able to ask for that care, using alcohol to cope and then feeling hopeless," she said.

Read DTN's special issue about mental health, "Rays of Hope Shedding Light on Rural Mental Health Challenges," at https://www.dtnpf.com/…

Todd Neeley can be reached at [email protected]

Follow him on social platform X @DTNeeley

(c) Copyright 2024 DTN, LLC. All rights reserved.

research studies on stress

Todd Neeley

Todd Neeley

  • Privacy Policy
  • Terms of Use
  • © 2024 DTN, all rights reserved. "DTN" and the degree symbol logo are trademarks of DTN. All other trademarks are the properties of their respective owners.
  • Magazine Home
  • Columns Home

MyDTN ™

  • MyDTN ™
  • The Progressive Farmer

Recommended Browsers:

  • Internet Explorer 10 or above
  • iPad 2 or above
  • iPhone 4 or above
  • Google Chrome for Android
  • Join Our Community
  • Start your free MyDTN demo

Please correct the following errors and try again:

  • Remember Me

News and features

Smell of human stress affects dogs’ emotions leading them to make more pessimistic choices.

research studies on stress

Press release issued: 22 July 2024

Dogs experience emotional contagion from the smell of human stress, leading them to make more ‘pessimistic’ choices, new research finds. The University of Bristol-led study, published in Scientific Reports today [22 July], is the first to test how human stress odours affect dogs' learning and emotional state.

Evidence in humans suggests that the smell of a stressed person subconsciously affects the emotions and choices made by others around them. Bristol Veterinary School researchers wanted to find out whether dogs also experience changes in their learning and emotional state in response to human stress or relaxation odours.

The team used a test of ‘optimism’ or ‘pessimism’ in animals, which is based on findings that ‘optimistic’ or ‘pessimistic’ choices by people indicate positive or negative emotions, respectively.

The researchers recruited 18 dog-owner partnerships to take part in a series of trials with different human smells present. During the trials, dogs were trained that when a food bowl was placed in one location, it contained a treat, but when placed in another location, it was empty.  Once a dog learned the difference between these bowl locations, they were faster to approach the location with a treat than the empty location.  Researchers then tested how quickly the dog would approach new, ambiguous bowl locations positioned between the original two.

A quick approach reflected ‘optimism’ about food being present in these ambiguous locations – a marker of a positive emotional state – whilst a slow approach indicated ‘pessimism’ and negative emotion. These trials were repeated whilst each dog was exposed to either no odour or the odours of sweat and breath samples from humans in either a stressed (arithmetic test) or relaxed (listening to soundscapes) state.

Researchers discovered that the stress smell made dogs slower to approach the ambiguous bowl location nearest the trained location of the empty bowl. An effect that was not seen with the relaxed smell.  These findings suggest that the stress smell may have increased the dogs’ expectations that this new location contained no food, similar to the nearby empty bowl location.

Researchers suggest this ‘pessimistic’ response reflects a negative emotional state and could possibly be a way for the dog to conserve energy and avoid disappointment.

The team also found that dogs continued to improve their learning about the presence or absence of food in the two trained bowl locations and that they improved faster when the stress smell was present.

Dr Nicola Rooney,  Senior Lecturer in Wildlife and Conservation at Bristol Veterinary School and the paper’s lead author explained: “Understanding how human stress affects dogs' wellbeing is an important consideration for dogs in kennels and when training companion dogs and dogs for working roles such as assistance dogs.

“Dog owners know how attuned their pets are to their emotions, but here we show that even the odour of a stressed, unfamiliar human affects a dog’s emotional state, perception of rewards, and ability to learn. Working dog handlers often describe stress travelling down the lead, but we’ve also shown it can also travel through the air.”

Dr Zoe Parr-Cortes, PhD student at Bristol Veterinary School and primary researcher on the project expressed her thanks to everyone involved in the study, especially all the participants and dog owners who took part in the research.

‘Does the odour of human stress or relaxation affect dogs’ cognitive bias? By Z. Parr-Cortes, N. J. Rooney et al. in Scientific Reports [open access].

Further information

Study information

The 18 dogs included in this study comprised ranged from eight months to ten years old.  Breeds consisted of two Springer spaniels; two Cocker spaniels; two Labrador Retrievers; two Braque d’Auvergne; one Whippet; one Golden Retriever; one Miniature Poodle and seven mixed breed dogs.  Eight dogs were registered as teaching dogs at the University of Bristol.

Study at the Bristol Veterinary School The University of Bristol offer a number of postgraduate courses including the MSc Global Wildlife Health and Conservation and a range of undergraduate degrees .

Based at Bristol's  Langford Campus , Bristol Veterinary School boasts first-class clinical facilities and encompasses a small animal hospital, a dairy farm, diagnostic laboratories, and farm animal, small animal and equine practices.

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings

Preview improvements coming to the PMC website in October 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • Health Psychol Open
  • v.7(2); Jul-Dec 2020

Best practices for stress measurement: How to measure psychological stress in health research

Despite the strong evidence linking psychological stress to disease risk, health researchers often fail to include psychological stress in models of health. One reason for this is the incorrect perception that the construct of psychological stress is too vague and broad to accurately measure. This article describes best practices in stress measurement, detailing which dimensions of stressor exposures and stress responses to capture, and how. We describe when to use psychological versus physiological indicators of stress. It is crucial that researchers across disciplines utilize the latest methods for measuring and describing psychological stress in order to build a cumulative science.

Introduction

Epidemiological studies confirm that both experiencing a greater number of stressful events and reporting high perceived stress over long periods of time are associated with worse mental and physical health, and mortality ( Epel et al., 2018 ). The association between greater stressor exposure and increased disease risk has been replicated with many different types of stressor exposures (e.g. discrimination, caregiving, work stress) and a range of aging-related health outcomes (e.g. cardiovascular disease, metabolic syndrome, mortality). The mechanistic pathways underlying these associations have also been detailed ( Boyce, 2015 ; McEwen, 2015 ; Miller et al., 2009 ). Despite this compelling evidence, however, health researchers often measure stress using unvalidated measures or select a single type of stress to measure, thus either missing entirely or underestimating the role stress plays in predicting disease onset or progression.

One of the main reasons for the lack of sophisticated measurement and inclusion of psychological stress in health models may be the incorrect assumption that stress is too broad and nebulous of a construct to accurately measure. It is true that psychological scientists too often fail to specify what they mean when using the term “stress” or other variants such as “stressor,” “acute stress,” “stress response,” and “stress biomarker.” Social and behavioral scientists tend to use the term loosely, often failing to define it clearly in a manuscript and using it to refer to a range of experiences, from living in poverty to giving a public speech to current negative mood. Kagan (2006) pointed out this lack of specificity, providing a fair critique of the state of the literature. The lack of specificity in language, however, does not represent a true lack of specificity in theoretical or methodological approaches. Although psychological stress researchers have made great strides in differentiating different forms of stress in recent decades, the problem is rather that the language used in journal articles has not always accurately reflected these advancements—and these advancements have been kept within a small, specialized subset of researchers. Thus, the purpose of this article is to provide health researchers across disciplines with a useful update on best practices for measuring stress and offer suggested language for how to describe stress-related constructs with more granular language.

Fundamentals of stress measurement

The term “stress” is an umbrella term representing experiences in which the environmental demands of a situation outweigh the individual’s perceived psychological and physiological ability to cope with it effectively ( Cohen et al., 2016 ). One important distinction in studying stress is to differentiate between exposures to stressful events and the responses to these events. Stressful events or “stressors” are discrete events that can be objectively rated as having the potential to alter or disrupt typical psychological functioning, such as losing your job or getting divorced. Stress responses are the cognitive, emotional, and biological reactions that these stressful events evoke.

Measuring stressor exposures versus stress responses

Stressor exposures can be measured with self-report questionnaires such as a life events checklist, assessed by an interviewer, or objectively determined based on proximity to an event (e.g. living in NYC during the September 11 terrorist attacks). The Life Events and Difficulties Schedule (LEDS; Brown and Harris, 1978 ) is a structured interview protocol that is considered the gold standard for assessing stressor exposure across someone’s lifetime. This interview protocol is time intensive in both the data collection and data processing stages. To streamline the process of capturing stressor exposures across the life span, a computer-assisted methodology was developed (e.g. The Stress and Adversity Inventory [STRAIN]; Slavich and Shields, 2018 ). In both the LEDS and the STRAIN, participants are asked whether they have experienced a range of stressful life events at any point in their life. For each endorsed stressor, they are asked follow-up questions to provide greater context about the experience (e.g. how old were you when it happened, how long did it go on for, how stressful or threatening was it). The LEDS requires a trained interviewer to administer the measure, while the STRAIN can be completed either by an interviewer or by participants themselves. The LEDS also relies on blind raters to score the severity of a stressor using this contextual information, while the STRAIN relies on the participants reporting of event severity. The STRAIN’s automated structure of follow-up questions allows the respondent to complete the interview much more quickly than the LEDS and reduces data processing time. Both measures provide a comprehensive assessment of stressor exposures across the lifespan, and use different methods to determine the severity of these experiences.

An individual’s response to the stressor sometimes matters more than mere exposure to it, particularly when it comes to the impact of the stressor on physical health. For example, caregiving for a family member with a debilitating illness is often considered a chronic stressor because of the constant physical and emotional demands. There is a significant amount of research examining the impact of being a dementia caregiver, in particular, given the large increase in the number of family dementia caregivers as the population ages in the United States. In fact, the Alzheimer’s Association estimated in 2018 that there were over 16 million family caregivers providing an estimated 18.5 billion hours of care to people with Alzheimer’s or other dementias ( Alzheimer’s Association, 2019 ). Empirical evidence has shown that family caregivers of Alzheimer’s patients have worse physical and mental health compared to age-matched non-caregivers ( Kiecolt-Glaser et al., 1987 ; Vitaliano et al., 2003 ). However, not every caregiver’s health is damaged by their caregiving role ( Roth et al., 2015 ). This may be because the negative impact of caregiving is caused by individuals’ subjective response to the caregiving situation, not from the mere exposure of being a caregiver. Thus, a better predictor of health decline would be the degree to which caregivers report high levels of psychological burden from their caregiving role. Empirical evidence supports this perspective; for example, Alzheimer’s caregivers who reported emotional distress or physical strain from caregiving had 63 percent greater mortality than caregivers who reported no distress ( Schulz et al., 1999 ).

Stress responses can be measured with self-report measures, behavioral coding, or via physiological measurements. These responses include emotions, cognitions, behaviors, and physiological responses instigated by the stressful stimuli. One of the simplest ways to measure stress responses is through self-reports of perceived stress related to a specific stressor or to one’s life circumstances ( Cohen et al., 1983 ). For example, the Perceived Stress Scale is a 10-item self-report measure that captures an individual’s perception of how overwhelmed they are by their current life circumstances. Responses to acute stressors have traditionally been studied in controlled laboratory settings in order to capture responses that unfold within minutes of stressor exposure (e.g. emotional and physiological reactivity to an acute stress task). A commonly used acute stress paradigm is the Trier Social Stress Test (TSST), a standardized laboratory stress task in which participants give a speech and perform mental arithmetic in front of judges ( Kirschbaum et al., 1993 ). The TSST reliably evokes an acute stress response for the majority of participants. Outside of the laboratory, new technology has enhanced our ability to capture real-time stress responses in daily life using mobile phones and wearables, which many researchers are now doing. Considering the impact of both stressor exposure and stress responses on health may improve the prediction of health outcomes, as many models of stress propose that the stress response mediates the effect of stress exposures on health outcomes ( McEwen, 1998 ; Wheaton et al., 2013 ).

Selecting stress measures

Due to constraints on participant burden and other considerations, difficult choices about which type of stress to measure need to be made by researchers. Common types of psychological stress measured using self-report questionnaires in adult samples are major life events, traumatic events, early life stress exposure, and current chronic or perceived stress in various domains (i.e. loneliness, marital discord, experiences of discrimination, work stress, financial strain, neighborhood safety and cohesion, and current perceived stress). The choice of which type of stressor exposure to measure depends on what is most relevant to the study population, the specific research question, and the hypothesized mechanisms linking that stress type to the outcome of interest. To begin the selection, consider first what is the most relevant stress type(s), given the sample’s demographic makeup. For example, measures that capture religious persecution or combat exposure would be particularly important for a sample living in a conflict zone, while the amount of overwhelm related to being a parent (parenting stress) may be most relevant for a sample of mothers caring for their child who has an autism spectrum disorder. In both cases, it would also be important to measure types of stressors that may not be directly related to the circumstances—such as levels of loneliness and financial strain. Capturing a range of stressor types reduces the likelihood that the individual’s psychological and social distress is underestimated.

Stressor and stress response characteristics

In addition to identifying stressor type(s) of interest, there are several key measurement considerations when choosing specific measures of stress to include in studies or analyzing existing stress measure data. These considerations include characteristics of the stressor or response (e.g. timescale, types of stressor response) as well as measurement characteristics (e.g. life stage of exposure and measurement assessment window). We briefly describe these aspects below (see Epel et al., 2018 for further discussion).

Timescale of the stressor

Stressors generally take place along the following timescales: chronic stressors, life events, daily events/hassles, and acute stress. Table 1 provides definitions for each of these timescales. It is important to note that naturalistic experiences of stress rarely fall neatly into one category. For example, death of a loved one is often considered a major life event but, depending on the cause of death, may also be considered a chronic stressor, such as if the family member was sick for years or months before the death. Similarly, arguments with a spouse may be considered an acute stressor, but if they happen every day they may be considered chronic. There is a significant amount of gray area between categories. Researchers should first make a thoughtful attempt to pick the category that best aligns with the stressor and with the way that stressor type has been described in past research, and then describe the exposure with as much specificity as possible.

Types of stress by timescale.

Type of stressDefinitionRelevance for health
Chronic stressChronic stressors are prolonged threatening or challenging circumstances that disrupt daily life and continue for an extended period of time (minimum of one month).People under the chronic stress are at greater risk of chronic illness, mortality, and accelerated biological aging ( ; ; ).
Life eventsLife events are time-limited and episodic events that involve significant adjustment to one’s current life pattern, such as getting fired, being in a car crash, or the death of a loved one. Some life events can be positive (e.g. getting married, moving to a new place), and some become chronic (e.g. disability caused by car crash).Exposure to more stressful life events is linked with poorer mental health in addition to development and progression of cardiovascular disease, as well as mortality due to cardiovascular disease and cancer ( ; ; ).
Traumatic life eventsTraumatic life events are a subclass of life events in which one’s physical and/or psychological safety is threatened.Experiencing a greater number of traumatic events across the life course is consistently associated with worse health and mortality ( ; ; ; ).
Daily hassles (i.e. daily stressors)Interruptions or difficulties that happen frequently in daily life such as minor arguments, traffic, or work overload, and that can build up overtime to create persistent frustration or overwhelm.Greater emotional responses to these daily hassles are associated with worse mental and physical health ( ; ; ; ).
Acute stressShort-term, event-based exposures to threatening or challenging stimuli that evoke a psychological and/or physiological stress response, such as giving a public speech.Greater cardiovascular reactivity to acute stressors has been prospectively associated with increased risk of cardiovascular disease ( ; ; ).

Types of stress response

Responses to stressor exposures provide additional useful information beyond measuring stressor exposure alone. Stress responses include psychological, behavioral, cognitive, and physiological reactions related to the stressor exposure that can occur before, during, or after the exposure. Psychological stress responses include specific emotions triggered by the stressor, as well as efforts to regulate that emotion ( Gross, 2002 ). Behavioral responses include coping behaviors such as smoking or seeking social support. Cognitive responses include appraisals of the exposure (e.g. as a threat versus challenge; Blascovich and Mendes, 2010 ) and perseverative cognitions (e.g. rumination Brosschot et al., 2005 ). Physiological responses include immune, autonomic, neuroendocrine, and neural changes related to stressor exposure. Further details about the various stress responses deserve more attention than can be described here ( Epel et al., 2018 ). As a part of selecting stress measures, researchers should identify the type of stress response that is most relevant for their research question and sample. Often, studies will assess multiple types of stress responses simultaneously.

Additional characteristics of the stressor

There are additional stressor exposure attributes that can be described and captured to thoroughly assess the exposure. These include, but are not limited to, duration, severity, controllability, life domain, the target of the stressor (e.g. self, close other), and the potential of the stressor to elicit specific harmful emotional responses (e.g. social status threat). Lack of control, social status threat, and stressor severity have been identified as potent attributes that predict worse outcomes across a range of stressor types and scenarios.

Characteristics of stress measurement

Life stage during stressor exposure.

In addition to the timescale of the stressor, another important characteristic of stressor exposure is the developmental or life stage during which the stressor occurs. Knowing the person’s age during the exposure informs hypotheses about which psychological and biological processes the stressor may have impacted. This is because developing systems are more open to environmental cues and are thus more likely to be impacted by stress exposure. “Sensitive periods” are specific time points in the life course during which physiological systems are maximally influenced by external environmental factors, and thus stressor exposure can have a particularly strong influence on development ( Knudsen, 2004 ). Sensitive periods during which stress may have the greatest effect are likely: prenatal ( Van Den Bergh et al., 2005 ; Weinstock, 2001 ), before age 5 ( Zeanah et al., 2011 ), during puberty ( Fuhrmann et al., 2015 ), entry into parenthood ( Saxbe et al., 2018 ), and during menopause ( Gordon et al., 2015 ). Identifying and measuring stress during sensitive periods could greatly increase our understanding of who is at risk for the negative effects of stress, the mechanistic pathways linking stress exposure to health decline, and where and how to focus intervention efforts.

Measurement assessment window

The window of measurement is also essential to consider to avoid measurement error and improve specificity in hypotheses. Measures can ask about stressors and stress responses across a wide range of time frames, such as in the present moment, over the course of that day, the past week, the past month, the past year, in childhood, or across the entire lifespan. For example, there are fundamental differences in a measure that ask participants to report on stress exposure in the past month versus across their lifespan. The latency between stressor exposure and measurement is crucial, as retrospective autobiographical reports are prone to bias and error, especially when there have been years or decades since the exposure in question ( Bradburn et al., 1987 ; Hardt and Rutter, 2004 ). In addition to the latency between exposure and measurement, several other factors can impact the accuracy of retrospective reports, such as mental state at the time of recall and the emotional salience of a given memory ( Shiffman et al., 2008 ). This may lead to overestimating the frequency of emotionally salient stressors and underestimating the frequency of more mundane, daily stressors ( Bradburn et al., 1987 ; Shiffman et al., 2008 ). For these reasons, it can be beneficial to measure stressor exposure and responses in close proximity to their occurrence whenever possible.

The experimental studies examining acute stressor exposure and responses, there are additional considerations with the measurement assessment window. Because the timing of stressor exposure is controlled, researchers can begin measuring psychological, behavioral, and physiological states prior to the stressor exposure and continue measuring throughout and after exposure. By measuring responses before, during, and after exposure, researchers can access (and predict) anticipation of and recovery from the stressor exposure.

Summary of steps for selecting stress measures

There are of course numerous considerations for selecting the appropriate stress measure for your study. In sum, researchers should identify the type or types of stress that are most relevant to their research question and sample. Stress measure selection should then be refined based on characteristics of the stressor and/or stress response that the researcher intends to measure, such as the timescale, the type(s) of stress responses the researcher is interested in, and other attributes of the stressor (e.g. duration, severity, controllability). Selection of stress measures should also account for measurement characteristics, such as the life stage during stressor exposure and the measurement assessment window (e.g. framing of questions, timing of assessment relative to occurrence of the stressor).

Beyond these stress-specific considerations, researchers should also follow general best practices for measure selection. For example, validated scales should be used when available. The Stress Measurement Network Toolbox provides a resource for validated measures of different types of stress that has beeen curated by experts ( https://stressmeasurement.org ). Measures should also be selected based on the uniqueness of the sample, and hile validated scales are preferred, some samples or exposures may require researchers to develop a new scale or modify an existing scale to fit their needs. These practical steps for selecting a stress measure are summarized in Table 2 .

Summary of steps for choosing appropriate stress measures.

Steps for choosing an appropriate measure of psychological stress.
1. Determine the type(s) of stress you intend to capture based on your research question and the uniqueness of your sample.
2. Determine the timescale of the stressor exposure and how you will capture objective exposure. For the exposure variable, in particular, you may need to develop your own measure based on the uniqueness of your sample.
3. Identify which types of stress responses you are able to assess in your study design (e.g. psychological, behavioral, cognitive, physiological).
4. Determine the life stage in which the stressor occurs and choose a measure appropriate for that particular life stage.
5. Identify additional characteristics of the stressor that are important to capture (e.g. severity, controllability, target of the stressor) and how these will be assessed (e.g. objective reviewer, participant report, a priori study design).
6. Consider your measurement assessment window and select measures that are specific to the time frame of exposure and/or response you intend to capture.
7. Look for well-validated scales that capture these aspects. It is common to use multiple scales to capture different aspects of the stress exposure and stress response, and the range of stress types that might be relevant for your sample. The Stress Measurement Network Toolbox provides validated and curated stress measures ( ).

Compelling evidence linking stress to physical health

The types of stress that have the most consistent and compelling relationships with disease risk and mortality are acute stress reactivity, early life stress, work or occupational stress, and social isolation/loneliness. A comprehensive review of these literatures is outside the scope of the present article; however, the following section highlights foundational studies linking these stress types physical health, with a particular emphasis on cardiovascular disease (because it is the leading cause of death in developed countries) and mortality. Effect sizes are included where possible, as are references to reviews and meta-analyses for further reading. Of note, we do not review the literature here on the impact of cumulative life stress (aggregate number of stressor exposures and/or intensity of stress responses over one’s life course). Despite initial compelling work on the impact of cumulative life stress on cardiovascular disease outcomes, this area of research is still in its infancy, with a need for measurement approaches to be unified across research studies to allow for building of a collective science ( Albert et al., 2013 ; Slopen et al., 2018 ).

Research on acute stress reactivity and physical health

Decades of research have shown that heightened cardiovascular reactivity and delayed recovery to acute stressors are prospectively associated with increased cardiovascular disease risk ( Brosschot et al., 2005 ; Chida and Steptoe, 2010 ; Steptoe and Marmot, 2005 ). One of the earliest studies in this area was a longitudinal study of healthy adult men (age 45–55; n  = 279) in which those classified as “hyper-reactors” (defined as > 20 mmHg increase in diastolic blood pressure to the cold pressor acute stress task) were 2.4 times more likely to have a myocardial infarction or die from cardiovascular disease in the following 20 years than men who showed a rise of < 20 mmHg ( Keys et al., 1971 ). Cortisol and inflammatory responses to acute stressors have also been shown to prospectively predict incident hypertension ( Hamer and Steptoe, 2012 ; Steptoe et al., 2016 ). Heightened reactions and prolonged recovery time periods may be driven by perseverative cognitions before (worrying) and after (rumination) stressor exposure ( Brosschot et al., 2005 , 2006 ). Despite the evidence linking reactivity to disease outcomes, the clinical meaningfulness of these associations is still debated ( Treiber et al., 2003 ). Importantly, a blunted response to an acutely stressful situation (sometimes termed a “hyporeactive response”), is also linked to worse health ( Carroll et al., 2017 ). For example, in a sample of 725 healthy adults from the Dutch Famine Birth Cohort Study, decreased cardiovascular and/or cortisol response to the acute stressor was associated with obesity, risk of becoming obese, depressive symptoms, anxiety, and poor self-rated and functional health ( De Rooij, 2013 ). In addition, there are several other reactivity patterns that have been hypothesized to represent maladaptive response profiles such as lack of habituation when exposed to repeated stressors of the same kind (see McEwen, 1998 ). Thus, the clinical meaningfulness of different stress reactivity profiles is largely debated.

Research on early life stress and physical health

The evidence linking early life stress to increased adult disease risk and mortality is strong. A foundational study in this area, the Adverse Childhood Experiences (ACE) Study, included nearly 10,000 adults and demonstrated that a greater number of self-reported retrospective adverse childhood experiences (e.g. physical abuse, living with an alcohol-dependent adult, witnessing violence) was positively associated in a graded relationship with the presence of ischemic heart disease, cancer, chronic lung disease, skeletal fractures, and liver disease, after controlling for demographic factors ( Felitti et al., 1998 ). Convincingly, reporting seven or more ACE was associated with three times the likelihood of heart disease compared to reporting no ACE ( Dong et al., 2004 ). These findings have been so compelling that significant changes in clinical and educational settings have been undertaken in recent years to recognize the role that early trauma has on current and future cognitive, socio-emotional, and behavioral outcomes for both children and adults.

Research on work stress and physical health

Epidemiological studies consistently demonstrate associations between high work stress and worse physical and mental health. One of the most widely studied models of work stress is job strain, which is a combination of high demands (workload and intensity) and low control ( Karasek, 1979 ). Decades of research has linked high job strain to anxiety and depression, increased blood pressure (BP), cardiovascular events, and metabolic syndrome ( Chandola et al., 2006 ; Landsbergis et al., 2013 ; Madsen et al., 2017 ; Nyberg et al., 2013 ). An analysis of the Whitehall II study cohort found that chronic work stress was associated with coronary heart disease (CHD) risk, with the associations strongest in participants under 50 (RR = 1.68, 95% CI 1.17–2.42). Other components of work stress, such as effort-reward imbalance, also predict cardiovascular disease risk ( Dragano et al., 2017 ).

Research on social isolation, loneliness, and physical health

A meta-analysis of decades of work on social isolation and loneliness found that being socially isolated, lonely, and/or living alone corresponded to an average of 29 percent, 26 percent, and 32 percent increased likelihood of mortality ( Holt-Lunstad et al., 2015 ). The mortality risk for the most socially isolated adults in the National Health and Nutritional Examination Survey (hazard ratio (HZ) = 1.62 for men, HZ = 1.75) was found to be comparable to the risk of smoking (HR = 1.72 for men, HZ = 1.86) and having high BP (HR = 1.16 for men, HR = 1.32 for women) ( Pantell et al., 2013 ). These strong relationships suggest that meaningful connection with others is an essential component of health and well-being. Several short measures have been created to capture this important social determinant of health, including a validated three-item measure of loneliness ( Hughes et al., 2004 ).

Biological pathways from stress to disease

There are numerous plausible biological pathways linking stress to cardiovascular disease, with most of the current evidence pointing to stress-related alterations in the immune, autonomic, and neuroendocrine systems. The brain networks that orchestrate stress-induced changes in these peripheral systems have also been identified ( Gianaros and Wager, 2015 ; Gianaros and Jennings, 2018 ), and can be described as the systems related to threat processing, safety processing, and social cognition ( Muscatell and Eisenberger, 2012 ). One widely accepted stress-disease model is the “wear and tear” hypothesis ( Charles et al., 2013 ; McEwen, 1998 ; Selye, 1956 ). This hypothesis is centered on the postulation that prolonged or repeated stress prematurely depletes the finite amount of “adaptational energy” of the organism, decreasing the body’s ability to successfully adapt to environmental challenges ( Selye, 1956 ). In this model, stressful events cause stress responses that involve activation of physiologic systems. In the short term, mobilizing physiological resources to respond to a discrete event or threat is adaptive. In the long term, however, frequent and repeated mobilization of these resources wears down these response systems and maladaptive patterns appear ( McEwen, 1998 ). The “wear and tear” hypothesis is theoretically compelling, but currently lacks definitive empirical support. This is because we do not currently have data that demonstrates the slow degradation of multiple physiological systems over decades in humans, an effort that requires tremendous investment. Instead, most studies have chosen one or maybe two physiological systems to measure to try to capture degradation or maladaptive responses to stressors, thus providing support, but not direct evidence for the “wear and tear” hypothesis. Other potential pathways include stress-related changes in endothelial function, elevated chronic inflammation, metabolic dysfunction, changes in DNA repair, changes in gene expression, and telomere shortening. These are all exciting areas of research, some of which fit in to the “wear and tear” hypothesis (e.g. telomere shortening; Epel et al., 2004 ) and others that suggest alternate processes (e.g. biological embedding of early experiences; Miller et al., 2011 ). These pathways are relevant for numerous chronic diseases beyond cardiovascular disease.

Associations between stress and immune system functioning are especially relevant given that the major diseases of aging in the United States are mediated, in part, through the immune system. The top three leading causes of death in the United States—cardiovascular disease, cancer, and chronic lower respiratory disease—all share the common thread of being characterized by elevated chronic inflammation ( Aghasafari et al., 2019 ; Golia et al., 2014 ; Grivennikov et al., 2010 ). Because of this common thread, chronic systemic inflammation has become a recent focus of health research. Stress exposure has been examined extensively as a predictor of increased systemic inflammation. Indeed, elevated systemic inflammation has been found in those experiencing chronic stress like caregivers ( Gouin et al., 2008 ), immediately after a stressful life event like death of a loved one ( Cohen et al., 2015 ), historical stress like childhood adversity ( Slopen et al., 2010 , 2012 ), daily stress ( Chiang et al., 2012 ), and in response to lab-based stress tasks ( Marsland et al., 2017 ). A short-term inflammatory response to stress is thought to be adaptive because it involves recruiting immune cells to the site of a real or potential injury in order to heal wounds resulting from stressor exposure. However, when there is no wound to heal, as is the case with psychosocial stressor exposure, repeated or exaggerated inflammatory responses may cause long-term damage and contribute to disease processes ( Black and Garbutt, 2002 ; Miller et al., 2002 ; Rohleder, 2014 ).

Is there an “objective” way to measure stress?

Stress and health researchers have searched for many years for a single biological indicator that someone is “under stress.” However, there is no single stress-specific biomarker. This is likely because acute stress is not the only state that evokes reliable biological changes (e.g. increased heart rate and BP). Other non-acute stress states, such as feeling excited, focusing attention on non-negative affect inducing stimuli, or exercising, also trigger biological responses that are similar to those evoked by negative affect inducing acute stressors like increased heart rate and blood pressure. This is even true for what is often termed the “stress hormone,” cortisol—not all cortisol increases are triggered by increases in psychological stress responses, nor does every experience that people perceive as “stressful” cause cortisol to rise ( Dickerson and Kemeny, 2004 ).

While measuring stress-related biomarkers may not provide a perfect indicator of whether someone is under stress or not, there are still compelling reasons to include these biomarkers in research studies of stress and health. Stress-related biomarkers are objectively measured biological indicators of physiological processes that are either implicated in the pathway from stress to disease or serve as a marker of that process. In typical models of the stress-health relationship, the stressful event (X) leads to a biological change (Y) that then leads to the disease state or related outcome (Z). Stress-related biomarkers can be the variable inserted in any component (X, Y, or Z) of this model; examples of the stress-related biomarker in each part of this basic model are shown in Figure 1 . In example A, the biomarker serves as a mediator, or a part of the causal pathway between a stressor and a health outcome. In example B, the biomarker serves as a predictor of stress-related psychosocial and behavioral processes that ultimately impact health outcomes. In example C, the biomarker serves as an outcome of psychological and physiological responses to a traumatic stressor. The way a biomarker is conceptualized (e.g. as a mediator, predictor, or outcome) depends on the research question and study methods. As such, choosing a stress-related biomarker to include in a study depends on the design of the study and the outcomes of interest. Table 3 provides further tools for how to choose the appropriate biomarker. It is also important to keep in mind that a biomarker may not be needed to answer a research question, despite the desire for a seemingly “objective” indicator of stress or stress reduction.

An external file that holds a picture, illustration, etc.
Object name is 10.1177_2055102920933072-fig1.jpg

Examples of how stress-related biomarkers can be modeled as either the predictor, the mediator, or the outcome in research studies.

Essential questions for following best practices in choosing an appropriate stress-related biomarker.

Questions to answer to help identify the right stress-related biomarker for your study:
1. What are the plausible biological pathways linking my stress predictors to my health outcome? The first step is to identify which physiological system is the likely candidate that is related to the health outcome of interest and that previous evidence has linked to stress or stress-related psychological processes.
2. What is the window of time that the stressor can plausibly have its impact for? If the stress response is short, is there a plausible reason it would have long-lasting impacts?
3. Is there a biomarker that captures functioning of the pathway identified in Question 1, and that reflects the appropriate timeline (Question 2)?
4. Is this biomarker associated with any end disease states relevant for my population of interest?
5. If you are proposing to use this biomarker for an intervention study, is the biomarker sensitive enough that it can change in the proposed intervention period window? Is it stable enough that the control condition would remain relatively stable during the intervention period? Would the expected change in the intervention group be clinically meaningful?
6. Are you able to collect the biomarker specimen well enough that is worth the subject burden and research cost? For example, while drawing blood is often the best way to capture many biomarkers, it is more invasive and requires more wet lab capacity than collecting saliva samples.
7. Is this biomarker needed to answer my research question or can this question be answered with a self-report or task-based measure? Biomarkers may not be needed despite initial excitement and desire to include a potentially “objective” indicator of stress or stress reduction.
8. For studies examining an acute stress response, what is the expected pattern of response? Complicating biomarker selection is that there is limited empirical evidence that identifies what a “bad” or “good” physiological acute stress response pattern is. This is because stress exposures take many forms, and thus the most adaptive response depends on a myriad of immediate contextual factors, such as what the goal of the arousal is.

One area of research that requires particularly careful consideration of biomarker selection is when biomarkers are used as an outcome in psychosocial intervention trials. The scientific community is often eager to find an objective biological indicator that a psychosocial intervention can improve health; this is typically done by measuring improvement in a biomarker from pre- to post-intervention. There has been a trend in recent years toward using changes in biomarkers as an indicator of an intervention’s success, rather than relying on subjective psychological reports of well-being. This approach is problematic for several reasons, including variability in baseline biomarker profiles, unknown reliability of biomarker assessment over time, unknown stability of these changes, and lack of evidence for the long-term impact of small changes in stress-related biomarkers on disease risk. Therefore, biomarkers should not replace self-report, behavioral, and cognitive outcomes as primary outcomes in psychosocial intervention trials aimed at reducing stress or related goals.

Variability in exposures and responses

Despite stress exposure being an inevitable part of life, not everyone develops stress-related illnesses at the same speed. One primary reason for this is that stress exposures are not distributed evenly across social groups. Women, young adults, members of racial-ethnic minority groups, divorced and widowed persons, and poor and working-class individuals report greater chronic stress and cumulative stress exposure across their lives ( Thoits, 2010 ). In addition, recent research has demonstrated that both psychological and physiological stress responses vary remarkably within and between people. While the physiological systems that are activated in response to a stressor are generally universal and non-specific as initially proposed by one of the founders of the field of stress, Hans Selye (1956) , the pattern of these responses vary considerably in terms of the degree of the system’s activation and how long the systems are activated for. Individual-level differences and environmental contexts interact to influence the psychological and physiological stress response trajectories. These include socioeconomic and cultural factors, genetic and developmental factors, historical and current stressors, stable protective factors, and health behaviors. A model integrating these different levels of experience is presented by our group in detail in Epel et al. (2018) and reprinted here with permission ( Figure 2 ).

An external file that holds a picture, illustration, etc.
Object name is 10.1177_2055102920933072-fig2.jpg

Transdisciplinary model of psychological stress: Integrating contextual, historical, habitual, and acute stress processes.

Figure 2 presents a transdisciplinary model that describes psychological stress as encompassing as a set of interrelated processes. The figure illustrates that stressors are experienced within the context of a person’s life, represented by the contextual factors in the blue triangle. These contextual factors include individual-level characteristics such as personality and demographics, the environment in which one lives, current and past stressor exposures, and protective factors—all of which combine to determine the baseline allostatic state of physiological regulation, and the lens through which stressors are perceived and assigned meaning. Contextual factors and habitual processes together influence psychological and physiological responses to acute and daily stressors. These responses, if dysregulated, are thought to lead to allostatic load and ultimately biological aging and early disease. Reprinted from Frontiers in Neuroendocrinology ( Epel et al., 2018 ).

Advanced statistical models can be used to examine variability in stress responses (both psychological and physiological) within and between people ( Bryk and Raudenbush, 1987 ; McArdle and Epstein, 1987 ). Within-person variability in stress responses means that a person’s response to a stressor within one life domain (e.g. work) does not necessarily predict how they will respond to a stressor within another life domain (e.g. family). Between-person variability means that different people respond to the same stressor in a variety of ways. As an example of variability in psychological stress responses, in a sample of 1,532 healthy adults from the Changing Lives of Older Couples prospective study, psychological responses to the death of one’s spouse took on four discrete trajectories (e.g. chronic grief, chronic depression, temporary depression, resilient), suggesting that there is not one universal pattern for spousal grief ( Galatzer-Levy and Bonanno, 2012 ). Cortisol can be used as an example of variability in physiological stress responses Cortisol generally increases in response to laboratory-based acute stress tasks if they are uncontrollable and characterized by social-evaluative threat ( Dickerson and Kemeny, 2004 ), such as the TSST described earlier ( Kirschbaum et al., 1993 ). However, around 30 percent of people do not mount a cortisol response, and there is tremendous variability in the size of the response. Individual-level predictors of this variability include age, gender, sex steroid levels, smoking, coffee, and alcohol consumption ( Kudielka et al., 2009 ). Interestingly, these differences are not driven by differences in the emotional responses to the task as acute stressors are not strongly correlated to the physiological responses. In a review of 49 acute stress studies, only 25 percent reported a significant correlation between the two emotional and physiological responses ( Campbell and Ehlert, 2012 ).

Empirical evidence supports a strong relationship between psychological stress and disease development. These studies may be underestimating the impact of stressor exposure and the stress response on health, given that measuring these constructs has been challenging and limited. Recent work in the stress field has identified important aspects of psychological stress to capture in order to fully test the role that psychological stress plays in predicting disease; these include capturing the specific type(s) of stressor exposure, a wide range of psychological, cognitive, behavioral, and physiological responses to the exposure, and contextual and individual-level factors that moderate the impact of the exposure and response. In this article, we identified ways for researchers to improve the language specificity when describing stress measures and offered guidance on how to choose the appropriate stress measure. We encourage the adoption of more precise language when writing about stress in academic papers, more careful selection of stress measures, with a focus on validated measures when possible, and theoretically driven integration of mechanistic pathways linking stress to health outcomes. The ultimate goal of having sophisticated research on the relationship between stress, health, and well-being is to develop evidence-based ways to help people thrive in our stress-filled world.

Acknowledgments

Members of the Stress Measurement Network provided essential input on the thoughts presented here, and we thank them for their contribution.

Conflict of Interests: The author(s) declared no potential conflicts of interest with respect to the research, authorship and/or publication of this article.

Funding: The author(s) disclosed receipt of the following financial support for the research, authorship, and/or publication of this article: This work was funded by the National Institute on Aging of the National Institutes of Health [R24AG048024; K01AG057859].

An external file that holds a picture, illustration, etc.
Object name is 10.1177_2055102920933072-img1.jpg

  • Aghasafari P, George U, Pidaparti R. (2019) A review of inflammatory mechanism in airway diseases . Inflammation Research 68 : 59–74. [ PubMed ] [ Google Scholar ]
  • Albert MA, Slopen N, Williams DR. (2013) Cumulative psychological stress and cardiovascular disease risk: A focused review with consideration of black-white disparities . Current Cardiovascular Risk Reports 7 : 318–325. [ Google Scholar ]
  • Almeida DM. (2005) Resilience and vunerability to daily stressors assessed via diary methods . Current Directions in Psychological Science 14 ( 2 ): 64–68. [ Google Scholar ]
  • Alzheimer’s Association (2019) 2019 Alzheimer’s disease facts and figures . Alzheimer’s & Dementia 15 ( 3 ): 321–387. [ PubMed ] [ Google Scholar ]
  • Black PH, Garbutt LD. (2002) Stress, inflammation and cardiovascular disease . Journal of Psychosomatic Research 52 : 1–23. [ PubMed ] [ Google Scholar ]
  • Blascovich JJ, Mendes WB. (2010) Social psychophysiology and embodiment . In: Fiske ST, Gilbert DT, Lindzey G. (eds) Handbook of Social Psychology . Hoboken, NJ: John Wiley & Sons, pp. 194–227. [ Google Scholar ]
  • Boyce WT. (2015) Epigenomics and the unheralded convergence of the biological and social sciences . In: Kaplan R, Spittel M, David DH. (eds) Population Health: Behavioral and Social Science Insights . Rockville, MD: Agency for Healthcare Research and Quality and Office of Behavioral and Social Sciences Research, National Institutes of Health, pp. 219–232. [ Google Scholar ]
  • Bradburn N, Rips L, Shevell S. (1987) Answering autobiographical questions: The impact of memory and inference on surveys . Science 236 ( 4798 ): 157–161. [ PubMed ] [ Google Scholar ]
  • Brosschot JF, Gerin W, Thayer JF. (2006) The perseverative cognition hypothesis: A review of worry, prolonged stress-related physiological activation, and health . Journal of Psychosomatic Research 60 : 113–124. [ PubMed ] [ Google Scholar ]
  • Brosschot JF, Pieper S, Thayer JF. (2005) Expanding stress theory: Prolonged activation and perseverative cognition . Psychoneuroendocrinology 30 ( 10 ): 1043–1049. [ PubMed ] [ Google Scholar ]
  • Brown G, Harris T. (1978) Social Origins of Depression: A Study of Psychiatric Disorders in Women . New York: The Free Press. [ Google Scholar ]
  • Bryk AS, Raudenbush SW. (1987) Application of Hierarchical Linear Models to Assessing Change . Psychological Bulletin 101 : 147–158. [ Google Scholar ]
  • Campbell J, Ehlert U. (2012) Acute psychosocial stress: Does the emotional stress response correspond with physiological responses? Psychoneuroendocrinology 37 ( 8 ): 1111–1134. [ PubMed ] [ Google Scholar ]
  • Carroll D, Ginty AT, Whittaker AC, et al. (2017) The behavioural, cognitive, and neural corollaries of blunted cardiovascular and cortisol reactions to acute psychological stress . Neuroscience and Biobehavioral Reviews 77 : 74–86. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Chandola T, Brunner E, Marmot M. (2006) Chronic stress at work and the metabolic syndrome: Prospective study . BMJ: British Medical Journal 332 : 521–525. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Charles ST, Piazza JR, Mogle J, et al. (2013) The wear and tear of daily stressors on mental health . Psychological Science 24 ( 5 ): 733–741. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Chiang JJ, Eisenberger NI, Seeman TE, et al. (2012) Negative and competitive social interactions are related to heightened proinflammatory cytokine activity . Proceedings of the National Academy of Sciences of the United States of America 109 ( 6 ): 1878–1882. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Chiang JJ, Turiano NA, Mroczek DK, et al. (2018) Affective reactivity to daily stress and 20-year mortality risk in adults with chronic illness: Findings from the National Study of daily experiences . Health Psychology 37 ( 2 ): 170–178. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Chida Y, Steptoe A. (2010) Greater cardiovascular responses to laboratory mental stress are associated with poor subsequent cardiovascular risk status: A meta-analysis of prospective evidence . Hypertension 55 ( 4 ): 1026–1032. [ PubMed ] [ Google Scholar ]
  • Chida Y, Hamer M, Wardle J, et al. (2008) Do stress-related psychosocial factors contribute to cancer incidence and survival? Nature Clinical Practice Oncology 5 ( 8 ): 466–475. [ PubMed ] [ Google Scholar ]
  • Cohen M, Granger S, Fuller-Thomson E. (2015) The association between bereavement and biomarkers of inflammation . Behavioral Medicine 41 ( 2 ): 49–59. [ PubMed ] [ Google Scholar ]
  • Cohen S, Gianaros P, Manuck S. (2016) A stage model of stress and disease . Perspectives on Psychological Science 11 ( 4 ): 456–463. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Cohen S, Janicki-Deverts D, Miller GE. (2007) Psychological stress and disease . Journal of the American Medical Association 298 : 1685–1687. [ PubMed ] [ Google Scholar ]
  • Cohen S, Kamarck T, Mermelstein R. (1983) A global measure of perceived stress . Journal of Health and Social Behavior 24 ( 4 ): 385–396. [ PubMed ] [ Google Scholar ]
  • De Rooij SR. (2013) Blunted cardiovascular and cortisol reactivity to acute psychological stress: A summary of results from the Dutch Famine Birth Cohort Study . International Journal of Psychophysiology 90 ( 1 ): 21–27. [ PubMed ] [ Google Scholar ]
  • Dickerson SS, Kemeny ME. (2004) Acute stressors and cortisol responses: A theoretical integration and synthesis of laboratory research . Psychological Bulletin 130 ( 3 ): 355–391. [ PubMed ] [ Google Scholar ]
  • Dong M, Giles WH, Felitti VJ, et al. (2004) Insights into causal pathways for ischemic heart disease: Adverse childhood experiences study . Circulation 110 : 1761–1766. [ PubMed ] [ Google Scholar ]
  • Dragano N, Siegrist J, Nyberg ST, et al. (2017) Effort-reward imbalance at work and incident coronary heart disease: A multicohort study of 90,164 individuals . Epidemiology 28 ( 4 ): 619–626. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Epel E, Blackburn EH, Lin J, et al. (2004) Accelerated telomere shortening in response to life stress . Proceedings of the National Academy of Sciences of the United States of America 101 ( 49 ): 17312–17315. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Epel E, Crosswell A, Mayer S, et al. (2018) More than a feeling: A unified view of stress measurement for population science . Frontiers in Neuroendocrinology 49 : 146–169. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Felitti VJ, Anda RF, Nordenberg D, et al. (1998) Household dysfunction to many of the leading causes of death in adults the Adverse Childhood Experiences (ACE) Study . American Journal of Preventive Medicine 14 ( 4 ): 245–258. [ PubMed ] [ Google Scholar ]
  • Fuhrmann D, Knoll LJ, Blakemore SJ. (2015) Adolescence as a Sensitive Period of Brain Development . Trends in Cognitive Sciences 19 : 558–566. [ PubMed ] [ Google Scholar ]
  • Galatzer-Levy IR, Bonanno GA. (2012) Beyond normality in the study of bereavement: Heterogeneity in depression outcomes following loss in older adults . Social Science & Medicine 74 ( 12 ): 1987–1994. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Gawronski KAB, Kim ES, Miller LE. (2014) Potentially traumatic events and serious life stressors are prospectively associated with frequency of doctor visits and overnight hospital visits . Journal of Psychosomatic Research 77 ( 2 ): 90–96. [ PubMed ] [ Google Scholar ]
  • Gianaros P, Wager T. (2015) Brain-body pathways linking psychological stress and physical health . Current Directions in Psychological Science 24 ( 4 ): 313–321. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Gianaros PJ, Jennings JR. (2018) Host in the machine: A neurobiological perspective on psychological stress and cardiovascular disease . American Psychologist 73 ( 8 ): 1031–1044. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Golia E, Limongelli G, Natale F, et al. (2014) Inflammation and cardiovascular disease: From pathogenesis to therapeutic target . Current Atherosclerosis Reports 16 ( 9 ): 435. [ PubMed ] [ Google Scholar ]
  • Gordon JL, Girdler SS, Meltzer-Brody SE, et al. (2015) Ovarian hormone fluctuation, neurosteroids, and HPA axis dysregulation in perimenopausal depression: A novel heuristic model . American Journal of Psychiatry 172 ( 3 ): 227–236. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Gouin J-P, Hantsoo L, Kiecolt-Glaser J. (2008) Immune dysregulation and chronic stress among older adults: A review . Neuroimmunomodulation 15 ( 4–6 ): 251–259. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Grivennikov SI, Greten FR, Karin M. (2010) Immunity, inflammation, and cancer . Cell 140 ( 6 ): 883–899. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Gross J. (2002) Emotion regulation: Affective, cognitive, and social consequences . Psychophysiology 39 : 281–291. [ PubMed ] [ Google Scholar ]
  • Hamer M, Steptoe A. (2012) Cortisol responses to mental stress and incident hypertension in healthy men and women . The Journal of Clinical Endocrinology and Metabolism 97 ( 1 ): E29–E34. [ PubMed ] [ Google Scholar ]
  • Hardt J, Rutter M. (2004) Validity of adult retrospective reports of adverse childhood experiences: Review of the evidence . Journal of Child Psychology and Psychiatry, and Allied Disciplines 45 ( 2 ): 260–273. [ PubMed ] [ Google Scholar ]
  • Holt-Lunstad J, Smith TB, Baker M, et al. (2015) Loneliness and social isolation as risk factors for mortality . Perspectives on Psychological Science 10 ( 2 ): 227–237. [ PubMed ] [ Google Scholar ]
  • Hughes ME, Waite LJ, Hawkley LC, et al. (2004) A short scale for measuring loneliness in large surveys: Results from two population-based studies . Research on Aging 26 ( 6 ): 655–672. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Kagan J. (2006) An overly permissive extension . Perspetives on Psychological Science 11 ( 4 ): 442–450. [ PubMed ] [ Google Scholar ]
  • Karasek R. (1979) Job demands, job decision latitude, and mental strain: Implications for job redesign . Administrative Science Quarterly 24 ( 2 ): 285–308. [ Google Scholar ]
  • Keyes KM, McLaughlin KA, Demmer RT, et al. (2013) Potentially traumatic events and the risk of six physical health conditions in a population-based sample . Depression and Anxiety 30 ( 5 ): 451–460. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Keys A, Longstreet Taylor H, Blackburn H, et al. (1971) Mortality and coronary heart disease among men studied for 23 years . Archives of Internal Medicine 128 ( 2 ): 201–214. [ PubMed ] [ Google Scholar ]
  • Kiecolt-Glaser J, Glaser R, Shuttleworth EC, et al. (1987) Chronic stress and immunity in family caregivers of Alzheimer’s disease victims . Psychosomatic Medicine 49 ( 5 ): 523–535. [ PubMed ] [ Google Scholar ]
  • Kirschbaum C, Pirke KM, Hellhammer DH. (1993) The ‘Trier social stress test’: A tool for investigating psychobiological stress responses in a laboratory setting . Neuropsychobiology 28 : 76–81. [ PubMed ] [ Google Scholar ]
  • Knudsen EI. (2004) Sensitive periods in the development of the brain and behavior . Journal of Cognitive Neuroscience 16 ( 8 ): 1412–1425. [ PubMed ] [ Google Scholar ]
  • Krause N, Shaw BA, Cairney J. (2004) A descriptive epidemiology of lifetime trauma and the physical health status of older adults . Psychology and Aging 19 ( 4 ): 637–648. [ PubMed ] [ Google Scholar ]
  • Kudielka BM, Hellhammer DH, Wüst S. (2009) Why do we respond so differently? Reviewing determinants of human salivary cortisol responses to challenge . Psychoneuroendo-crinology 34 ( 1 ): 2–18. [ PubMed ] [ Google Scholar ]
  • Landsbergis PA, Dobson M, Koutsouras G, et al. (2013) Job strain and ambulatory blood pressure: A meta-analysis and systematic review . American Journal of Public Health 103 ( 3 ): e61–e71. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • McArdle JJ, Epstein D. (1987) Latent growth curves within developmental structural equation models . Child Development 58 ( 1 ): 110–133. [ PubMed ] [ Google Scholar ]
  • McEwen BS. (1998) Stress, adaptation, and disease: Allostasis and allostatic load . Annals of the New York Academy of Sciences 840 : 33–44. [ PubMed ] [ Google Scholar ]
  • McEwen BS. (2015) The brain on stress: How behavior and the social environment “get under the skin.” In: Kaplan R, Spittel M, David DH. (eds) Population Health: Behavioral and Social Science Insights . Rockville, MD: Agency for Healthcare Research and Quality and Office of Behavioral and Social Sciences Research, National Institutes of Health, pp. 233–250. [ Google Scholar ]
  • Madsen IEH, Nyberg ST, Magnusson Hanson LL, et al. (2017) Job strain as a risk factor for clinical depression: Systematic review and meta-analysis with additional individual participant data . Psychological Medicine 47 ( 8 ): 1342–1356. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Marsland AL, Walsh C, Lockwood K, et al. (2017) The effects of acute psychological stress on circulating and stimulated inflammatory markers: A systematic review and meta-analysis . Brain, Behavior, and Immunity 64 : 208–219. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Miller GE, Chen E, Cole SW. (2009) Health psychology: Developing biologically plausible models linking the social world and physical health . Annual Review of Psychology 60 ( 1 ): 501–524. [ PubMed ] [ Google Scholar ]
  • Miller GE, Chen E, Parker KJ. (2011) Psychological stress in childhood and susceptibility to the chronic diseases of aging: Moving toward a model of behavioral and biological mechanisms . Psychological Bulletin 137 ( 6 ): 959–997. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Miller GE, Cohen S, Ritchey AK. (2002) Chronic psychological stress and the regulation of pro-inflammatory cytokines: A glucocorticoid-resistance model . Health Psychology 21 ( 6 ): 531–541. [ PubMed ] [ Google Scholar ]
  • Muscatell KA, Eisenberger NI. (2012) A social neuroscience perspective on stress and health . Social and Personality Psychology Compass 6 : 890–904. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Nyberg ST, Fransson EI, Heikkilä K, et al. (2013) Job strain and cardiovascular disease risk factors: Meta-analysis of individual-participant data from 47,000 men and women . PLoS ONE 8 ( 6 ): e67323. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Pantell M, Rehkopf D, Jutte D, et al. (2013) Social isolation: A predictor of mortality comparable to traditional clinical risk factors . American Journal of Public Health 103 ( 11 ): 2056–2062. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Rohleder N. (2014) Stimulation of systemic low-grade inflammation by psychosocial stress . Psychosomatic Medicine 76 ( 3 ): 181–189. [ PubMed ] [ Google Scholar ]
  • Rosengren A, Wilhelmsen L, Orth-Gomér K. (2004) Coronary disease in relation to social support and social class in Swedish men: A 15 year follow-up in the study of men born in 1933 . European Heart Journal 25 ( 1 ): 56–63. [ PubMed ] [ Google Scholar ]
  • Roth DL, Fredman L, Haley WE. (2015) Informal caregiving and its impact on health: A reappraisal from population-based studies . Gerontologist 55 ( 2 ): 309–319. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Saxbe D, Rossin-Slater M, Goldenberg D. (2018) The transition to parenthood as a critical window for adult health . American Psychologist 73 ( 9 ): 1190–1200. [ PubMed ] [ Google Scholar ]
  • Schulz R, Beach SR. (1999) Caregiving as a risk factor for mortality: The caregiver health effects study . JAMA 282 ( 23 ): 2215–2219. [ PubMed ] [ Google Scholar ]
  • Selye H. (1956) The Stress of Life . New York: McGraw-Hill. [ Google Scholar ]
  • Shiffman S, Stone AA, Hufford MR. (2008) Ecological momentary assessment . Annual Review of Clinical Psychology 4 ( 1 ): 1–32. [ PubMed ] [ Google Scholar ]
  • Sin NL, Graham-Engeland JE, Ong AD, et al. (2015) Affective reactivity to daily stressors is associated with elevated inflammation . Health Psychology 34 ( 12 ): 1154–1165. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Slavich GM, Shields GS. (2018) Assessing lifetime stress exposure using the Stress and Adversity Inventory for Adults (Adult STRAIN): An overview and initial validation . Psychosomatic Medicine 80 : 17–27. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Slopen N, Koenen KC, Kubzansky LD. (2012) Childhood adversity and immune and inflammatory biomarkers associated with cardiovascular risk in youth: A systematic review . Brain, Behavior, and Immunity 26 ( 2 ): 239–250. [ PubMed ] [ Google Scholar ]
  • Slopen N, Lewis TT, Gruenewald TL, et al. (2010) Early life adversity and inflammation in African Americans and whites in the midlife in the United States survey . Psychosomatic Medicine 72 ( 7 ): 694–701. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Slopen N, Meyer C, Williams D. (2018) Cumulative stress and health . In: Ryff CD, Krueger R. (eds) The Oxford Handbook of Integrative Health Science . Oxford: Oxford University Press, pp. 75–93. [ Google Scholar ]
  • Steptoe A, Kivimäki M. (2013) Stress and cardiovascular disease: An update on current knowledge . Annual Review of Public Health 34 ( 1 ): 337–354. [ PubMed ] [ Google Scholar ]
  • Steptoe A, Marmot M. (2005) Impaired cardiovascular recovery following stress predicts 3-year increases in blood pressure . Journal of Hypertension 23 ( 3 ): 529–536. [ PubMed ] [ Google Scholar ]
  • Steptoe A, Kivimäki M, Lowe G, et al. (2016) Blood pressure and fibrinogen responses to mental stress as predictors of incident hypertension over an 8-year period . Annals of Behavioral Medicine 50 ( 6 ): 898–906. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Thoits P. (2010) Stress and health: Major findings and policy implications . Journal of Health and Social Behavior 51 ( S ): 41–53. [ PubMed ] [ Google Scholar ]
  • Treiber FA, Kamarck T, Schneiderman N, et al. (2003) Cardiovascular reactivity and development of preclinical and clinical disease states . Psychosomatic Medicine 65 : 46–62. [ PubMed ] [ Google Scholar ]
  • Van Den Bergh BRH, Mulder EJH, Mennes M, et al. (2005) Antenatal maternal anxiety and stress and the neurobehavioural development of the fetus and child: Links and possible mechanisms. A review . Neuroscience and Biobehavioral Reviews 29 : 237–258. [ PubMed ] [ Google Scholar ]
  • Vitaliano PP, Zhang J, Scanlan JM. (2003) Is Caregiving Hazardous to One’s Physical Health? A Meta-Analysis . Psychological Bulletin 129 : 946–972. [ PubMed ] [ Google Scholar ]
  • Weinstock M. (2001) Alterations induced by gestational stress in brain morphology and behaviour of the offspring . Progress in Neurobiology 65 : 427–451. [ PubMed ] [ Google Scholar ]
  • Wheaton B, Young M, Montazer S, et al. (2013) Social stress in the twenty-first century . In: Aneshensel CS. (ed.) Handbook of the Sociology of Mental Health (2nd edn). Cham: Springer, pp. 299–323. [ Google Scholar ]
  • Zeanah CH, Gunnar MR, McCall RB, et al. (2011) Sensitive periods . Monographs of the Society for Research in Child Development 76 ( 4 ): 147–162. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Open access
  • Published: 15 July 2024

Endoplasmic reticulum stress promotes hepatocellular carcinoma by modulating immunity: a study based on artificial neural networks and single-cell sequencing

  • Zhaorui Cheng   ORCID: orcid.org/0009-0008-3474-1161 1 ,
  • Shuangmei Li 1   na1 ,
  • Shujun Yang 1 ,
  • Huibao Long 1 ,
  • Haidong Wu 1 ,
  • Xuxiang Chen 1 ,
  • Xiaoping Cheng 2 &
  • Tong Wang   ORCID: orcid.org/0000-0003-4644-4179 1  

Journal of Translational Medicine volume  22 , Article number:  658 ( 2024 ) Cite this article

389 Accesses

Metrics details

Introduction

Hepatocellular carcinoma (HCC) is characterized by the complex pathogenesis, limited therapeutic methods, and poor prognosis. Endoplasmic reticulum stress (ERS) plays an important role in the development of HCC, therefore, we still need further study of molecular mechanism of HCC and ERS for early diagnosis and promising treatment targets.

The GEO datasets (GSE25097, GSE62232, and GSE65372) were integrated to identify differentially expressed genes related to HCC (ERSRGs). Random Forest (RF) and Support Vector Machine (SVM) machine learning techniques were applied to screen ERSRGs associated with endoplasmic reticulum stress, and an artificial neural network (ANN) diagnostic prediction model was constructed. The ESTIMATE algorithm was utilized to analyze the correlation between ERSRGs and the immune microenvironment. The potential therapeutic agents for ERSRGs were explored using the Drug Signature Database (DSigDB). The immunological landscape of the ERSRGs central gene PPP1R16A was assessed through single-cell sequencing and cell communication, and its biological function was validated using cytological experiments.

An ANN related to the ERS model was constructed based on SRPX, THBS4, CTH, PPP1R16A, CLGN, and THBS1. The area under the curve (AUC) of the model in the training set was 0.979, and the AUC values in three validation sets were 0.958, 0.936, and 0.970, respectively, indicating high reliability and effectiveness. Spearman correlation analysis suggests that the expression levels of ERSRGs are significantly correlated with immune cell infiltration and immune-related pathways, indicating their potential as important targets for immunotherapy. Mometasone was predicted to be the most promising treatment drug based on its highest binding score. Among the six ERSRGs, PPP1R16A had the highest mutation rate, predominantly copy number mutations, which may be the core gene of the ERSRGs model. Single-cell analysis and cell communication indicated that PPP1R16A is predominantly distributed in liver malignant parenchymal cells and may reshape the tumor microenvironment by enhancing macrophage migration inhibitory factor (MIF)/CD74 + CXCR4 signaling pathways. Functional experiments revealed that after siRNA knockdown, the expression of PPP1R16A was downregulated, which inhibited the proliferation, migration, and invasion capabilities of HCCLM3 and Hep3B cells in vitro.

The consensus of various machine learning algorithms and artificial intelligence neural networks has established a novel predictive model for the diagnosis of liver cancer associated with ERS. This study offers a new direction for the diagnosis and treatment of HCC.

As one of the most commonly primary malignancies, liver cancer has become one of the top five causes of cancer-related death around the world according to the World Health Organization [ 1 ]. Approximately 90% of liver cancer patients die from hepatocellular carcinoma (HCC) which is the most common pathological type [ 2 ]. Surgical treatment remains the most effective way to HCC. However, due to the insidious onset and rapid progression of HCC, patients frequently fail to avail themselves of the surgical option because of delayed medical consultation [ 3 ]. Moreover, HCC patients often face the daunting challenge of high prevalence in chemotherapy drug resistance, distant metastasis and recurrences, consequently resulting in an unfavorable prognosis [ 3 , 4 ]. Therefore, it is vital to deeply investigate the underlying mechanism of HCC occurrence and development, so that we can find new and promising targets for diagnosis and treatment of HCC patients.

Endoplasmic reticulum (ER) involved in lipid and carbohydrate metabolism and calcium strorage [ 5 , 6 ]. Moreover, as the largest and the most powerful organelle in eukaryotic cells, ER is also mainly responsible for the synthesis, transportation and folding of protein [ 5 , 6 ]. Endoplasmic reticulum stress (ERS) refers to the protein folding disorder in ER under pathological or physiological stimuli, such as activation of oncogenes, oxidative stress, hypoxia, and infection [ 7 , 8 ]. ERS regulate three main pathway of unfolded protein response (UPR), including PRKR-like ER kinase, activated transcription factor 6, and inositol requirement Enzyme 1 which alleviate the load of unfolded proteins load, and maintain cell homeostasis and function [ 9 ]. UPR pathways are activated in most cancer types because protein synthesis increases dramatically during the rapid proliferation of tumor cells [ 10 , 11 ]. As the initiating factor of UPR, ERS plays a crucial role in the therapy response and prognosis of cancer. At the beginning of chemotherapy, drugs cause deficiencies in nutrients and hypoxia of tumor cells, which lead to the ERS followed by UPR [ 12 , 13 ]. Once the UPR is activated, tumor cells release pro-survival components including cytokines, growth factors, and other factors, which induce cancer cell growth and proliferation and suppressing anti-tumor immune response [ 14 , 15 ], It is reported that when HCC mice were treated with the IRE1α-inhibitor, alleviation of tumor load and collagen accumulation were observed, which indicate that regulating ERS and UPR is an effective way to inhibit drug resistance to HCC.

In our study, machine learning techniques such as Random Forest (RF) and Support Vector Machine (SVM) algorithms were applied to screen for key genes associated with hepatocellular carcinoma (HCC). Subsequently, by integrating these genes, an artificial neural network was utilized to construct an ERS-related HCC diagnostic model. On the training set and three validation sets, the diagnostic model exhibited satisfactory predictive performance. We also conducted a comprehensive analysis of the expression levels, immune infiltration, methylation, and mutation status of ERSRGs. Our research offers a novel perspective on understanding the molecular mechanisms of HCC and identifies potential targets for developing new diagnostic and therapeutic strategies for HCC.

Data sources used for analysis

The author first integrated gene expression matrices from GSE25097, GSE62232, and GSE65372, analyzed the gene expression differences between normal and liver cancer tissues, and conducted functional enrichment analysis. By comparing the intersection of differentially expressed genes with genes related to endoplasmic reticulum stress, and employing two machine learning methods, six candidate biomarkers were identified, including SRPX, THBS4, CTH, PPP1R16A, CLGN, and THBS1. Based on these genes, an artificial neural network (ANN) algorithm was utilized to construct a diagnostic model. Subsequently, the diagnostic performance of these candidate genes was validated in three independent validation sets (GSE121248, GSE45267, and GSE84005). Moreover, molecular docking was employed to screen potential target drugs, and the immune cell infiltration rate, methylation level, and mutation rate of the marker genes were assessed. It was found that PPP1R16A exhibited a high copy mutation rate and was significantly correlated with the level of immune cell infiltration. To further identify PPP1R16A as a core gene in the endoplasmic reticulum stress model, single-cell sequencing and cell communication analyses were conducted to study its expression and distribution patterns in the tumor microenvironment. Finally, the biological function of the PPP1R16A gene was validated through in vitro experiments. The overall design of this study is illustrated in Fig.  1 .

figure 1

The overall flow of this study

Data collection and preprocessing

Transcriptome data and clinical information of HCC patients and normal tissue donors were from the GEO database ( https://www.ncbi.nlm.nih.gov/geo/ ). Three datasets were obtained, including GSE25097 (249 normal tissues and 268 tumor tissues), GSE62232 (10 normal tissues and 81 tumor tissues), and GSE65372 (15 normal tissues and 39 tumor tissues). These datasets were combined and removed repeating tissues using “sva” R package. A total of 662 samples were obtained, and the expression matrix of 14,738 genes was used as the training set. Besides, validation sets consisted of three datasets including GSE121248 (37 normal tissues and 70 tumor tissues), GSE45267 (39 normal tissues and 48 tumor tissues), and GSE84005 (38 normal tissues and 38 tumor tissues). All data is standardized and log-transformed by using the R “limma” software package for subsequent analysis [ 16 ].

The identification of differential expressed genes and ERSRGs

The “limma” R package was used to detect differential expressed genes (DEGs) between HCC and normal tissues in the training set with |log2 fold change (FC)|> 1.5 and adjusted p  < 0.05 as cutoff value. The volcano plot and heat map showing the differential expression of genes between HCC and normal tissue were made using the “ggplot2” and “heatmap” R packages. The Gene Ontology (GO) and Kyoto Encyclopedia of Genes and Genomes (KEGG) functional enrichment analysis were conducted among DEGs using “clustersProfiler”, “enrichplot”, “limma”, “ggplot2” and “org.Hs.eg.db” R package. The hallmark gene set “h.all.v7.4.symbols.gmt” was downloaded from MSigDB datasets ( https://www.gsea-msigdb.org/) 17 and used for gene set variance analysis (GSVA) analysis with the p  < 0.05 and false discovery rate (FDR) < 0.25. In addition, 15 hallmark gene sets including 312 ERSRGs were also downloaded from the MSigDB database.

The construction and validation of artificial neural network (ANN) prediction model using artificial intelligence algorithms

With “random Forest” R package, we established the RF model using fivefold cross-validation method to iterate on the variables’ number at each split and tree. When the number of branches was 125, we got the minimum residual error. We ranked genes according to Gini coefficient score and those genes with score > 20 were finally selected [ 18 , 19 ]. Using the “e1071” and “caret” R package, the SVM algorithm is applied to delete the SVM-generated feature vectors and identify the optimal variables. We fitted a linear SVM model, sorted the variables by their weights and eliminated the variables with low weights. Through the cross validation, the number of selected genes was determined when root mean square error is minimal [ 20 ]. ERSRG with RF and SVM algorithms, the ANN predictive model was established. The ANN model was constructed based on a multilayer perceptron network using R package “neuralnet” and “NeuralNetTools”. This ANN model include input layers, hidden layers, and output layers, and was tested using back-propagation algorithms. The first layer through input layers, neurons transmit the weighted data to neural groups, and then the hidden layers was applied to randomly select bias. Once the hidden nodes’ net sums is validated, the output responses were provided through transfer function [ 21 ]. In our study, six ERSRG were selected as the input nodes, as well as HCC and normal tissues were used as the output nodes. Besides, the predictive performance of ANN prognostic models was evaluated through the area under curve (AUC) of time-dependent receiver operating characteristics (ROC) curves using R package “pROC”.

Survival and methylation analysis of HCC patients

Survival analysis for six ERSRG was performed using the gene expression profiling interactive analysis (GEPIA) website ( http://gepia.cancer-pku.cn/index.html ). The Kaplan–Meier (K-M) survival analysis was performed to compare patient differences between high and low expression of ERSRG group in conjunction with the log-rank test. Obtain the methylation levels of the six ERSRGs from the UALCAN website ( https://ualcan.path.uab.edu/ )

Immune, co-mutation and genetic alterations analysis of six ERSRG in HCC patients

The infiltration of 29 type of immune cells and immune-related pathways were analyzed in HCC patients through ssGSEA analysis using the “GSVA” and “GSEABase” R package [ 22 , 23 ]. The relationship of ERSRG expression with immune cell infiltration and pathway enrichment was identified through Spearman analysis for coefficient calculation. Genomic data of six ERSRG including somatic mutations and DNA copy-number alterations was obtained from cBioPortal website ( https://www.cbioportal.org/ ). Besides, co-mutation analysis of six ERSRG was applied with “corplot” R package to explore the expression correlation among them.

Single-cell RNA-seq analysis

GSE149614, liver cancer single cell data set, which includes 10 liver cancer samples. Seurat package (version 4) is used for single cell data processing and analysis. Specifically, 3,4411 cells are obtained by preliminary screening according to the data of genes expressed in at least 3 cells and cells expressing at least 200 genes. Further, according to the secondary screening conditions of 500 < nfeature < 5000 and mitochondrial gene ratio < 10%, cell populations retained in each sample is shown in Table  1 .

Normalize data + find variable features + scale data function pipeline was used to standardize and normalize the data, and runpca function was used to calculate the top 50 principal components. According to the results of jackstraw and elbow plot, the top 20 principal components are the most appropriate. The find clusters function identifies 31 clusters with a resolution of 0.4 to 30, and umap performs dimensionality reduction clustering.

Cell-cell interaction analysis

The cell communication was analyzed using the R package ‘cell chat’. First, according to the expression of the protein phosphatase 1 regulatory subunit 16 A (PPP1R16A), the presence or absence of hepatocytes was divided into PPP1R16A positive and negative groups, and they were analyzed together with other cells. Additionally, the netAnalysis_compute Centrality function compared the outgoing signals and incoming signals among different cells to determine the core pathways mediating cell-cell interactions, and the hierarchy plot visualization was performed for some selected pathways. The netAnalysis_contribution function calculated the contribution of the receptor-ligand pairs in specific pathways and displayed the ligand-receptor pairs with the highest contribution.

Cell culture and transfection

The liver cancer cell lines HCCLM3, HepG2, Hep3B, and the normal liver cell line LO2 were all purchased from Shanghai Fuheng Cell Biology Co., Ltd. and cultured according to standard procedures. Lipid-mediated siRNA transfection was performed using the lipo3000 reagent (Invitrogen, USA) according to the siRNA product instructions (Ribio, China). The siRNA sequences targeting PPP1R16A are as follows: siRNA#1 - TGCCCGAAATGACCTGGAA; siRNA#2 - TGCGGCATCTATACTCCAA; siRNA#3 - CCAACATCAATGCCTGTGA.

Real-time quantitative qRT-PCR

Total RNA from HCC cells was extracted using the TRIzol reagent (Invitrogen) and reverse transcribed into cDNA using PrimeScript Reverse Transcriptase (Takara, Japan) before qRT-PCR. Quantitative PCR was performed using SYBR Premix EX TaqTM II (Takara, Japan) and the LightCycler 480 real-time PCR system (Roche, Shanghai, China). GAPDH was used as an endogenous control gene to normalize the expression of the target gene. Each sample was analyzed in triplicate. The thermal cycling program included holding for 10 s at 95 °C, 30 s at 60 °C, and 60 s at 72 °C. Then, melt curve data was collected. The primer sequences are shown in Table  2 .

Cell proliferation, migration and invasion assay

After transfecting cells with siRNA for 48 h, 3,000 cells per well were seeded into 96-well plates. Each well was contained 100 µl complete growth medium and 10µL CCK-8 was added and mixed in each well. After 2 h, the absorbance at 450 nm was measured. Cell migration and invasion were determined using a Transwell chamber. In the migration assay, cells (5*10 3 ) after 48 h of siRNA transfection were seeded into the upper chamber (Corning) of serum-free medium, and the lower chamber was filled with medium containing 20% FBS. After approximately 24 h, they were fixed with 4% paraformaldehyde for 20 min and stained with 0.1% crystal violet for 15 min. The invasion assay followed the same procedure as the migration assay, except that the lower chamber of the invasion experiment was coated with Matrigel (BD Biosciences).

Drug prediction and molecular docking

The drug prediction for the core genes of the endoplasmic reticulum stress model was conducted using the DSigDB database from Enrichr ( https://maayanlab.cloud/Enrichr/ ). Subsequently, the small molecules were docked with the aforementioned central targets using AutoDock Vina (Scripps Research, San Diego, CA). The docking results were evaluated and analyzed using the PLIP system ( https://plip-tool.biotec.tu-dresden.de/plip-web/plip/index ). Finally, the molecular docking (MD) results in two-dimensional structure were visualized using LIGPLOT software version 4.5.3, and the MD diagrams were generated using PyMOL. Protein structures were obtained from the PDB ( https://www.pdb.org/ ) or AlphaFold ( https://alphafold.com/ ), and drug-related data were retrieved from PubChem ( https://pubchem.ncbi.nlm.nih.gov/ ).

Cell scratch

Cells were seeded equably in 6-wll plate, and then transfected for 48 h. Approximately 5*10 5 cells were added to each well, and 10uL needles were used to scrape three horizontal lines on the surface of plate. Following this, the cells were cultured in a 2% serum medium for 24 h. Optical microscopy was used to collect images of cells at after 0 and 24 hours.

Statistical analysis

Version R 4.1.3 was used for all statistical studies. Differences between two groups were compared through Mann–Whitney test for non-normally distributed variables and unpaired t-test for normally distributed variables. Spearman analysis was used for correlation analysis and coefficient calculation. We analyzed the RT-qPCR results using paired t-tests and drew a Venn diagram using funrich software. P value < 0.05 was defined as statistically significant.

The identification of DEGs between HCC and normal tissues

We performed principal component analysis (PCA) of genes of datasets GSE25097, GSE62232 and GSE65372, under both non-batch remove and batch remove condition respectively, and results were presented in Fig.  2 A and B. The results show that we successfully removed batch effects from different data sets. The volcano plot in Fig.  2 C showed the 763 DEGs between HCC and normal tissues, of which 215 genes were up-regulated and 548 genes were down-regulated. The heatmap in Fig.  2 D also presented the expression level of DEGs between HCC and normal tissues.

figure 2

Differential gene expression analysis between liver cancer and normal tissues.( A ) Principal component analysis (PCA) of genes without batch removal for datasets including GSE25097, GSE62232, and GSE65372. ( B ) PCA of genes with batch removal for datasets including GSE25097, GSE62232, and GSE65372. ( C ) A volcano plot representing 763 differentially expressed genes (DEGs) between liver cancer tissues and normal tissues. ( D ) A heatmap showing the 763 DEGs between HCC and normal tissues

GO and KEGG as well as GESA enrichment analysis

The functional enrichment of DEGs between HCC and normal tissues was analyzed. The biological process in GO analysis of DEGs mainly included the alpha-amino acid metabolism, small molecule catabolism, carboxylic acid catabolism, organic acid catabolism, cellular amino acid metabolism. The cellular component in GO analysis was enriched in collagen-containing extracellular matrix, collagen trimers, blood microparticles, mitotic spindle and chromosomal regions. The molecular function in GO analysis was mainly involved in monooxygenase activity, iron ion binding, oxidoreductase activity acting on paired donors, with incorporation or reduction of molecular oxygen, heme binding, and steroid hydroxylase activity (Fig.  3 A-C). Besides, the KEGG analysis showed that the DEGs mainly contributed to tryptophan metabolism, complement and coagulation cascades, PPAR signaling pathway, cell cycle, and tyrosine metabolism in Fig.  3 D. Furthermore, the GSVA analysis of DEGs was conducted to explore the molecular pathway enrichment of DEGs. The results showed that DEGs were mainly abundant in hallmark MYC targets V1, hallmark E2F targets, hallmark inflammatory response, hallmark TNF-α signal via NF-κB and hallmark interferon-γ response (Fig.  3 E).

figure 3

Functional enrichment of DEGs between liver cancer tissues and normal tissues.( A - C ) Gene ontology (GO) analysis of DEGs, including molecular function (MF), cellular component (CC), and biological process. ( D ) Kyoto Encyclopedia of Genes and Genomes (KEGG) analysis of DEGs. ( E ) Gene Set Variation Analysis (GSVA) of DEGs.

The identification of ERS-related DEGs

The 11 ERS-related DEGs were obtained through the intersection of 312 ERSRGs and 763 DEGs in Fig.  4 A.A RF analysis further screened 11 ERSRGs. The residual graph in Fig.  4 B showed that the residual was the smallest when the number of branches was 125. As presented in Figs.  4 C and 11 ERSRGs were scored in this RF tree, and 7 of them with a score of > 20 were finally selected according to the descending Gini coefficient. Additionally, SVM analysis was also applied to screening the 11 ERSRGs. As depicted in Fig.  4 D, the root mean square error was the smallest when 6 of 11 ERSRGs were retained. Through the intersection of 7 ERSRGs in RF analysis and 6 ERSRGs in SVM analysis, 6 ERSRGs including SRPX, THBS4, CTH, PPP1R16A, CLGN, and THBS1 were finally determined for the construction of the ANN below (Fig.  4 E).

figure 4

Identification of ERS-related differentially expressed genes. (A) ERS-related genes ( n  = 312) were cross-referenced with DEGs ( n  = 763) to identify ERS-related differentially expressed genes (ERSRGs) ( n  = 11). ( B ) Residual plot for the selection of ERSRGs using the Random Forest (RF) algorithm. ( C ) The 11 selected ERSRGs are arranged in descending order according to the Gini coefficient. ( D ) Selection of ERSRGs using the Support Vector Machine (SVM) algorithm. ( E ) Identification of 6 ERSRGs through the intersection of genes selected by RF and SVM.

Construction and validation of ANN prediction model in HCC

The ANN was established including 6 neurons (SRPX, THBS4, CTH, PPP1R16A, CLGN, and THBS1) as the input layer and 2 neurons (HCC and normal tissues, respectively) as the output layer (Fig.  5 A). When the hidden layer included 5 neurons, the predictive performance of the model was the highest with the AUC value of 0.979 (95% CI: 0.968–0.989) in the training group (Fig.  5 B). The efficiency of this ANN predictive model was also assessed in the other 3 validation groups. The AUC value of the ANN model was 0.958 (95% CI: 0.914–0.992), 0.936 (95% CI: 0.852-1.000) and 0.970 (95% CI: 0.920-1.000) in GSE121248, GSE45267 and GSE84005, respectively (Fig.  5 C-E). Above all, this ANN prediction model based on the ER is a potentially powerful tool to predict HCC diagnosis.

figure 5

Construction and Validation of the Artificial Neural Network (ANN) Prediction Model for Liver Cancer. ( A ) The ANN comprises 6 neurons as the input layer, 2 neurons as the output layer, and 5 neurons as the hidden layer. ( B ) The AUC value of the ANN prediction model in the training set. ( C - E ) The AUC values of the ANN prediction model in the validation groups

The impact of endoplasmic reticulum stress core genes on immune infiltration deserves attention

The expression of six ERSRGs between HCC and normal tissues was deeply investigated in Fig.  6 A. The results showed that there were significantly higher expression levels of CLGN, PPP1R16A and THBS4 and lower expression levels of CTH, SRPX and THBS1 in HCC tissues compared with normal tissues, implying that CLGN, PPP1R16A and THBS4 as oncogenic genes played an important role in HCC development and progression while CTH, SRPX and THBS1 as onco-suppressor genes may play an inhibitory role in HCC. The interaction of 6 ERSRGs were further analyzed to identify the co-occurrence or mutually exclusive relationship of them in Fig.  6 B. There was an intense co-occurrence relationship between THBS1 and SRPX with a correlation coefficient of 0.67 as well as a mutually exclusive relationship between SRPX and THBS4 with a correlation coefficient of -0.53. In order to explore the underlying mechanism that these ERSRGs contributed to HCC occurrence and progression, the relationship of 6 ERSRGs expression levels with the immune cell infiltration and immune-related pathways was analyzed as presented in Fig.  6 C. We can conclude that THBS1 and SRPX were significantly positively associated with the infiltration of most immune cells and immune-related pathways such as neutrophils, T helper cells, tumor-infiltrating lymphocytes (TILs), cytokines and chemokines receptors (CCR), and para-inflammation etc. In contrast, THBS4, PPP1R16A and CLGN were significantly negatively associated with the infiltration of most of immune cells and immune related pathways such as neutrophils, T helper cells, TILs, CCR, checkpoint, T cell co-inhibition. Above findings further demonstrated that the ERSRGs can regulate tumor immune micro-environment to expose effects on HCC prognosis.Genes related to the TNF family molecules and chemotactic factors have been collected from previous literature. The ggcor and ggpplot2 packages were utilized to analyze and visualize the expression correlations between core genes and two types of immune activity molecules. The results show green squares representing positive correlations, purple squares indicating negative correlations, solid lines representing positive correlations, and dashed lines depicting negative correlations. Notably, CLGN, SRPX, and THBS1 exhibit positive correlations with most immune activity molecules, whereas PPP1R16A, CTH, and THBS4 show negative correlations with most immune activity molecules (Fig.  6 D). This conclusion is largely consistent with the results of the immune cells and processes depicted in the preceding figure.

figure 6

Expression and Immune infiltration Analysis of Six ERSRGs. ( A ) The expression of six ERSRGs between HCC and normal tissues. ( B ) The Co-mutation analysis of six ERSRGs. ( C ) The relationship of six ERSRGs expression levels with the immune cell infiltration and immune-related pathway. ( D )Correlation between the expression levels of 6 ERSRGs and immune-related molecules

Drug screening and molecular docking of characteristic genes

The results showed the top 20 drugs with the highest score values, among which L-cysteine had the highest binding score, while Mometasone had the strongest binding significance (Fig.  7 A). Therefore, we conducted further molecular docking model simulations of the target genes bound by these two drugs, and the results showed that THBS1 had the lowest binding energy with Mometasone compared to CTH, which was − 6.917 kcal/mol. Therefore, Mometasone may be the most suitable therapeutic drug for THBS1 (Fig.  7 B-D).

figure 7

Predict the top 20 candidate drugs for endoplasmic reticulum stress-related genes based on PubChem. ( A ) Predict the top 20 most significant candidate compounds for ERSRGs using the DSigDB database. ( B - G ) Molecular docking between THBS1 and Mometasone

The six ERSRGs-related survival, mutation and methylation analysis

It was depicted from the Kaplan–Meier (KM) curve that the HCC patients with low expression levels of CLGN and PPP1R16A had significantly longer overall survival than those with high expression levels of two genes (Fig.  8 A and C). By contrast, HCC patients with high expression levels of CTH statistically lived longer compared with those with low expression level of CTH (Fig.  8 B). However, there was no significant difference in the prognosis of HCC patients between high and low expression of SRPX, THBS1 and THBS4 which needed further validation (Fig.  8 D-F). In addition, gene mutation and copy number variation of 6 ERSRGs in 379 HCC patients from the cBioportal website was compared in Fig.  7 G. There was the highest mutation rate of PPP1R16A among 6 ERSRGs which was characterized by gene amplification. In addition to describing the survival and mutation status of core ERS genes, we also analyzed the methylation status of the promoters corresponding to these core genes. According to the calculation method on the UALCAN website ( https://ualcan.path.uab.edu/ ), we found that there were significant differences in the degree of promoter methylation between the control group and the liver cancer cell group for CLGN, CTH, PPP1R16A, and THBS1. Among them, the promoter methylation level of CLGN was stronger in the tumor group, while the methylation levels of the other three genes were significantly reduced in the tumor group, especially PPP1R16A (Fig.  8 H-M).

figure 8

Survival, mutation and methylation Analysis Related to ERSRGs. ( A - F ) Kaplan-Meier curves representing the differences in overall survival between groups with high and low expression levels of six ERSRGs. ( G ) Comparison of gene mutations and copy number variations among the six ERSRGs. ( H - M ) Calculation of methylation levels of six ERSRGs genes

Single-cell sequencing analysis of the immune landscape of the PPP1R16A gene

The UMAP diagram shows that all cells can be annotated into seven types of cells, including liver parenchymal cells, macrophages and so on (Fig.  9 A). The bubble chart shows the marker genes used in seven different annotation cell groups. Such as liver Parenchymal cells (CD24, MDK), Macrophages (CD68, CD163), Fibroblasts (PDGFRb), Endothelial cells (PECAM1), T/NK cells (CD3D, CD3E), Plasma cells (JSRP1), B cells (MS4A1) (Fig.  9 B). Single cell analysis showed that PPP1R16A was mainly expressed in malignant liver parenchymal cells (Fig.  9 C). Go enrichment analysis showed that the cells with high expression of PPP1R16A were mainly related to the process of lipid metabolism, such as cholesterol metabolism, fatty acid metabolism and so on (Fig.  9 D). KEGG enrichment analysis showed that PPP1R16A was related to glucose metabolism, lipid metabolism and PPAR signaling pathway. These results suggest that PPP1R16A is highly enriched in liver parenchyma, which may aggravate the progression of liver cancer by affecting metabolism related pathways (Fig.  9 E).To further explore the downstream pathways of PPP1R16A, we conducted a GSEA analysis of KEGG at the single-cell level. The differential analysis results of all genes were sorted by their logFC, and GSEA analysis was performed using the clusterProfiler package and GSVA package. The resulting KEGG analysis results were further clustered and visualized using the aPEAR package. The results showed that among all significant signaling pathways, seven pathways had clustered modular characteristics and were more core signaling pathways (Fig.  9 F).

figure 9

Single-cell RNA-seq Data Analysis (GSE149613). (A) UMAP plot of annotated cell types. ( B ) Bubble chart showing markers corresponding to different cell types. ( C ) UMAP diagram shows the expression of ppp1R16A ( D ) Go enrichment analysis, including BP, CC, MF. ( E )KEGG enrichment analysis; The color close to blue indicates a smaller p value, and the larger bubble table indicates that more differential genes are enriched in this pathway.( F )Analysis of downstream signaling pathways of GSEA at the single-cell level

Exploration of the impact of the PPP1R16A gene in the tumor microenvironment through cellular communication

Based on the expression of PPP1R16A in malignant liver parenchymal cells, we can divide the liver parenchymal cells into PPP1R16A_pos and PPP1R16A_neg. Using the Cellchat package to calculate the cell communication between these two types of cells and other cell types, we found that fibroblasts, endothelial cells, and PPP1R16A_pos cells are the core cells in the communication network, with fibroblasts being the cell with the strongest output signals. From the perspective of PPP1R16A_pos as the signal sender, it has more communication numbers and intensities with endothelial cells and macrophages (Fig.  10 A-B). PPP1R16A_pos cells had more communication with VTN, PARs, complement, CD46, PROS, CADM, GDF, CDH, and OCLN compared to other cells. When looking only at PPP1R16A_pos cells themselves, the strongest output signaling pathway was the macrophage migration inhibitory factor (MIF) pathway. In a lateral comparison, PPP1R16A_pos cells had more communication with MK, FN1, ANGPTL, THBS, PTN, CADM, EGF, CDH, and OCLN. When looking only at PPP1R16A_pos cells themselves, the strongest received signal was the COLLAGEN pathway (Fig.  10 C). We further studied the specific interactions between MIF and COLLAGE pathways. Hierarchical diagram shows that PPP1R16A_pos cells, as MIF signal senders, mainly send signals to T/NK cells, macrophages, B cells, and plasma cells (Fig.  10 D). As a COLLAGEN signal receiver, it mainly receives signals from endothelial cells and fibroblasts, of which fibroblasts received the most signals (Fig.  10 E). Further mining of the ligands and receptors that are most likely to play a role in the pathway, the highest pairing probability in the MIF pathway was between the MIF ligand and the CD74 + CXCR4 receptor (Fig.  10 F), and the highest probability of pairing in the COLLAGEN pathway was between the COL4A1 ligand and the SDC1 receptor (Fig.  10 G).

figure 10

Inference of cell–cell communications in TME. ( A - B ) A cell–cell communications between the identified cell types. ( C ) The incoming and outgoing signaling pathways of each cell type. ( D - G ) The hierarchical diagram displays the specific interaction between MIF and COLLAGE pathways

Upregulation of PPP1R16A in liver cancer and knockdown of PPP1R16A significantly suppresses the proliferative, invasive, and migratory abilities of HCC cells

By summarizing all the aforementioned findings, we have established that PPP1R16A was a pivotal gene associated with ERS and exhibiting a high copy number mutation. QRT-PCR experiments, showed that human HCC cell lines exhibited a higher expression of PPP1R16A compared to normal liver cells (Fig.  11 A). We simultaneously transfected siRNA into Hep3B and HCCLM3 cells, selecting the highly most effective siRNA-PPP1R16A#1 for subsequent experiments (Fig.  11 B and C). CCK8 experiments indicated that knockdown of PPP1R16A suppressed the proliferation capacity ability of Hep3B and HCCLM3 cells (Fig.  11 D). Meanwhile, scratch test and Transwell experiments demonstrated that knockdown of PPP1R16A inhibited the migration and invasion abilities of HCCLM3 and Hep3B cells. These observations suggested that PPP1R16A is a positive regulator of HCC cells (Fig.  11 E and F).

figure 11

Validation of PPP1R16A Expression Levels and Knockout of PPP1R16A Significantly Inhibits Proliferation, Invasion, and Migration Capabilities of HCC ( A )Expression levels of PPP1R16A mRNA in HCC cell lines. ( B - C ) Knockout efficiency of PPP1R16A mRNA in Hep3B and HCCLM3 cells. ( D ) CCK8 assays indicate that the knockdown of PPP1R16A inhibits the proliferation abilities of Hep3B and HCCLM3 cells. ( E - F ) Knockout of PPP1R16A inhibits the migration and invasion capabilities of Hep3B and HCCLM3 cells (* p  < 0.05, ** p  < 0.01, *** p  < 0.001, ns: not significant)

Hepatocellular carcinoma (HCC) is a type of tumor with poor prognosis, exhibiting a high incidence rate and mortality rate globally. It is a highly refractory disease, despite advancements in surgical and systemic treatments, the prognosis for HCC remains unsatisfactory. Early diagnosis and treatment can improve the survival rate of HCC patients. It is imperative to detect and identify the progression characteristics of the tumor at an early stage. Therefore, developing new treatment techniques and identifying new therapeutic targets is of critical importance.Currently, endoplasmic reticulum stress has garnered widespread attention from various cancer researchers. During endoplasmic reticulum stress, when cells are overly stimulated, the unfolded protein response (UPR) is triggered. This response influences the regulation of the endoplasmic reticulum (ER) balance through three distinct receptors: IRE1α, PERK, and ATF6. Research has shown that the activation of ERS affects tumor cell proliferation, invasion, metastasis, and promotes rapid tumor progression [ 24 , 25 ].However, the biological mechanisms underlying ERSRGs remain unclear, and their impact on HCC warrants further exploration。.

In this study, we conducted a comprehensive analysis of transcriptomic data between liver cancer and normal tissues. We identified differentially expressed genes (DEGs) between liver cancer and normal tissues, as well as the functions and molecular pathway enrichment of these DEGs. Based on the expression profile analysis of the training set, six endoplasmic reticulum stress-related genes (ERSRGs) were identified using the Random Forest (RF) and Support Vector Machine (SVM) algorithms. Subsequently, an artificial neural network (ANN) prediction model was constructed and demonstrated effective predictive performance. This model was further validated on three independent test sets, confirming its superior predictive capability. We also conducted an in-depth study on the association and function of these genes in tumorigenesis and immunomodulation.

The current (ERSRGs) encompass six potential genes (SRPX, THBS4, CTH, PPP1R16A, CLGN, and THBS1). In fact, previous studies have elucidated the significant roles of some of these genes in various tumors. Cystathionine γ-lyase, encoded by the CTH gene, plays a crucial role in the cysteine sulfur metabolism pathway. It catalyzes the generation of hydrogen sulfide (H 2 S), L-cysteine, α-ketobutyrate, and ammonia [ 26 ]. Several studies have indicated that aberrant activation of the CTH/H 2 S signaling pathway is closely linked to the occurrence and progression of HCC [ 27 ]. X-rays activate the p38 mitogen-activated protein kinase, which in turn activates the CTH/H 2 S signaling pathway, inducing epithelial-mesenchymal transition and promoting invasion of liver cancer cells [ 28 ]. Recent research has also reported that FOXC1, by regulating CTH, inhibits cysteine metabolism, increases reactive oxygen species levels, and promotes tumorigenesis. Overexpression of CTH significantly inhibits the proliferation, invasion, and metastasis of liver cancer cells induced by FOXC1 [ 29 ]. In contrast, CTH presents a potential therapeutic target when normally regulated in contrast to FOXC1. Furthermore, Sushi repeat-containing protein X-linked (SRPX) has been identified as a potential therapeutic target in HCC treatment. SRPX has been identified through mRNA expression network analysis, and has been shown to suppress cancer cell stemness [ 30 ]. SRPX also regulates the migration and invasion of ovarian cancer through the Ras homolog family member A signaling pathway [ 31 ]. Thrombospondin-1 (THBS1), known for inhibiting angiogenesis, has been studied for its potential as a therapeutic target [ 32 ]. THBS1 promotes the progression and development of various cancers by regulating angiogenesis and tumor vascular perfusion [ 33 ]. Additionally, THBS1 modulates innate and adaptive immune cells through the CD47 signaling molecules, thereby restricting anti-tumor immunity [ 34 ]. Overexpression of THBS4 promotes the proliferation and migration of liver cancer cells, participates in the regulation of epithelial-mesenchymal transition progression and interacts with members of the integrin family to modulate the FAK/PI3K/AKT pathway [ 35 ]. The miR-142 is highly correlated with THBS4 overexpression in HCC tissue samples, by regulating THBS4 expression in HCC cells [ 36 ], PPP1R16A encoded the membrane-associated subunit of protein phosphatase 1 which is located on the plasma membrane as a CAR-binding protein [ 37 ]. The area under the ROC curve for PPP1R6A in global and initial-stage tumors was 0.82 and 0.76, respectively, showing excellent sensitivity and specificity to define the diagnosis likelihood of endometrial carcinoma [ 38 ]. However, the role of PPP1R6A in HCC diagnosis and prognosis is rarely known and requires further exploration. A recent study reported that upregulation of CLGN in HCC is significantly related to poor prognosis, especially in advanced stages which might be regulated by miR-194-3p, thus providing a potentially therapeutic target and prognosis predictor in HCC [ 39 ].

In cellular communication analysis, we found that the expression levels of these ERSRGs were closely associated with immune cell infiltration and the activity of immune-related pathways. Single-cell sequencing revealed that the high expression of PPP1R16A in the liver parenchyma may be a trigger for high-copy mutations. Given that MIF acted as a macrophage stimulator, we speculate from cell communication results that PPP1R16A cells may promote macrophage aggregation through the MIF pathway, which inducing M2 polarization of liver cancer cells. Recent study suggests that novel ERSRGs signature could an independent prognostic factor for HCC [ 40 ]. ERS regulate immune levels by regulating myeloid cells, mainly macrophages which is related to tumor evasion of the immune response, and chemoresistance. ER-stressed HCC cells release exosomes, upregulate the expression of PD-L1 in macrophages, and consequently suppresses T-cell function. A higher density of infiltrated macrophages in the liver has been observed to be associated with enhanced tumor aggressiveness and unfavorable prognosis among patients with HCC [ 11 ].

In recent years, immune checkpoint inhibitors (ICIs) as a new therapeutic approach targeting T cells regulatory pathways, have much attention [ 41 , 42 ], and have great prospects in the field of anti-tumor therapy. Through the analysis of ssGSEA results, we found that THBS1 and SRPX showed significant positive correlations with immune cell infiltration, neutrophils, helper T cells, TILs, CCR, and inflammatory response-related pathways. In contrast, THBS4, PPP1R16A, and CLGN exhibited significant negative correlations with immune cell infiltration, neutrophils, helper T cells, TILs, CCR, immune checkpoints, T cell co-inhibition, and other immune-related pathways. These findings suggest that ERSRGs is closely related to the immune status of liver cancer, and offered a new research direction for the combination targeting of ERSRGs and ICIs in the treatment of liver cancer. Through combination therapy, there is potential to enhance anti-tumor immune responses and improve the prognosis of liver cancer.

We also found in cellular communication analysis that, simultaneously, fibroblasts, as essential factors promoting tumor metastasis, may reshape the tumor microenvironment by enhancing the collagen pathway and promoting collagen deposition to affect the function of PPP1R16A cells. These results further indicate that PPP1R16A may influence the prognosis of HCC by regulating the tumor immune microenvironment. Additionally, our experimental results suggest that knocking out PPP1R16A can inhibit the proliferation, invasion, and migration capabilities of HCC cells, indicating that PPP1R16A may be a crucial tumor-promoting factor. Cancer-associated fibroblasts produce collagen and change the extracellular matrix, which is an important mechanism of tumor metastasis. Modulating targeting specific signaling molecules responsible for crosstalk between Cancer‐associated fibroblasts and tumor cells is considered a promising approach to modulating HCC metastasis [ 43 , 44 ]. Our study indicated that PPP1R16A may be one of such potential targets.

However, it is important to note that there are some limitations that need further addressing and in-depth exploration. Firstly, Considering the bioinformatics analysis based on public cancer databases, it is crucial to further validate the diagnostic and predictive performance of ERSRG markers in large-scale and prospective clinical trials and assess their potential clinical applications. This will contribute to ensuring the reliability and reproducibility of the analysis results and provide a more solid foundation for the clinical application of ERSRGs in liver cancer patients. Secondly, Despite some cell experimental validation was involved, providing support for preliminary findings, further in vivo and in vitro experiments are needed to thoroughly investigate the functions of ERSRGs in HCC. This expanded experimental research will contribute to a more comprehensive and in-depth understanding of the exact mechanisms of action of these genes in the development and progression of HCC, and contribute to a more comprehensive understanding of the potential efficacy and mechanisms of ERSRGs in combination with ICIs in liver cancer. Therefore, future research directions should include broader experimental designs to more comprehensively and systematically reveal the role of ERSRGs in the biology of HCC biology.

Conclusions

In this study, the researchers integrated the six identified ERSRGs into an ANN prediction model based on RF and SVM algorithms. Furthermore, we further investigated the biological mechanisms, immune regulation, and genomic mutations associated of these six ERSRGs in the diagnosis of liver cancer. The comprehensive analysis of ERSRGs provides a powerful tool for the prognosis prediction and personalized treatment of liver cancer patients. The feature model based on ERSRGs holds promise as a crucial prediction and therapeutic decision support system in the field of liver cancer research.

Data availability

All data generated or analyzed during this study are included in this published article.

Abbreviations

Hepatocellular Carcinoma

Artificial Neural Network

Endoplasmic Reticulum Stress

Protein Phosphatase 1 Regulatory Subunit 16 A

Area Under the Curve

Endoplasmic Reticulum Stress Related Gene

Unfolded Protein Respons

Differential Expressed Genes

Gene Expression Omnibus

Gene Ontology

Kyoto Encyclopedia of Genes and Genomes

Gene Set Variance Analysis

Receiver Operating Characteristic

Uniform Manifold Approximation and Projection

Random Forest

Support Vector Machine

Gene Expression Profiling Interactive Analysis

Principal Components

Sushi Repeat-Containing Protein X-linked

Thrombospondin-1

Macrophage Migration Inhibitory Factor

Immune Checkpoint Inhibitors

Sung H, Ferlay J, Siegel RL, Laversanne M, Soerjomataram I, Jemal A, Bray F. Global Cancer statistics 2020: GLOBOCAN estimates of incidence and Mortality Worldwide for 36 cancers in 185 countries. CA Cancer J Clin. 2021;71:209–49.

Article   PubMed   Google Scholar  

Llovet JM, Kelley RK, Villanueva A, Singal AG, Pikarsky E, Roayaie S, Lencioni R, Koike K, Zucman-Rossi J, Finn RS. Hepatocellular carcinoma. Nat Rev Dis Primers. 2021;7:6.

Forner A, Reig M, Bruix J. Hepatocellular carcinoma. Lancet. 2018;391:1301–14.

Cronin KA, Scott S, Firth AU, Sung H, Henley SJ, Sherman RL, Siegel RL, Anderson RN, Kohler BA, Benard VB, Negoita S, Wiggins C, Cance WG, Jemal A. Annual report to the nation on the status of cancer, part 1: National cancer statistics. Cancer. 2022;128:4251–84.

Fagone P, Jackowski S. Membrane phospholipid synthesis and endoplasmic reticulum function. J Lipid Res. 2009;50:S311–316.

Article   PubMed   PubMed Central   Google Scholar  

Schwarz DS, Blower MD. The endoplasmic reticulum: structure, function and response to cellular signaling. Cell Mol Life Sci. 2016;73:79–94.

Article   CAS   PubMed   Google Scholar  

Ma Y, Hendershot LM. ER chaperone functions during normal and stress conditions. J Chem Neuroanat. 2004;28:51–65.

He J, Li G, Liu X, Ma L, Zhang P, Zhang J, Zheng S, Wang J, Liu J. Diagnostic and prognostic values of MANF expression in hepatocellular carcinoma. Biomed Res Int. 2020;2020:1936385.

Xu H, Tian Y, Tang D, Zou S, Liu G, Song J, Zhang G, Du X, Huang W, He B, Lin W, Jin L, Huang W, Yang J, Fu X. An endoplasmic reticulum stress-microRNA-26a feedback circuit in NAFLD. Hepatology. 2021;73:1327–45.

Pavlović N, Heindryckx F. Exploring the role of endoplasmic reticulum stress in hepatocellular carcinoma through mining of the human protein atlas. Biology (Basel). 2021;10:640.

PubMed   Google Scholar  

Pavlović N, Heindryckx F. Targeting ER stress in the hepatic tumor microenvironment. FEBS J. 2022;289:7163–76.

Kokott-Vuong A, Jung J, Fehr AT, Kirschfink N, Noristani R, Voigt A, Reich A, Schulz JB, Huber M, Habib P. Increased post-hypoxic oxidative stress and activation of the PERK branch of the UPR in trap1-deficient drosophila melanogaster is abrogated by metformin. Int J Mol Sci. 2021;22:11586.

Article   CAS   PubMed   PubMed Central   Google Scholar  

Pavlović N, Calitz C, Thanapirom K, Mazza G, Rombouts K, Gerwins P, Heindryckx F. Inhibiting IRE1α-endonuclease activity decreases tumor burden in a mouse model for hepatocellular carcinoma. Elife. 2020;9:e55865.

Hou J, Zhang H, Sun B, Karin M. The immunobiology of hepatocellular carcinoma in humans and mice: basic concepts and therapeutic implications. J Hepatol. 2020;72:167–82.

Tacke F. Targeting hepatic macrophages to treat liver diseases. J Hepatol. 2017;66:1300–12.

Ritchie ME, Phipson B, Wu D, Hu Y, Law CW, Shi W, Smyth GK. Limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res. 2015;43:e47.

Nayarisseri A, Khandelwal R, Tanwar P, Madhavi M, Sharma D, Thakur G, Speck-Planche A, Singh SK. Artificial intelligence, big data and machine learning approaches in precision medicine & drug discovery. Curr Drug Targets. 2021;22:631–55.

Guo L, Wang Z, Du Y, Mao J, Zhang J, Yu Z, Guo J, Zhao J, Zhou H, Wang H, Gu Y, Li Y. Random-forest algorithm based biomarkers in predicting prognosis in the patients with hepatocellular carcinoma. Cancer Cell Int. 2020;20:251.

Lin X, Yang F, Zhou L, Yin P, Kong H, Xing W, Lu X, Jia L, Wang Q, Xu G. A support vector machine-recursive feature elimination feature selection method based on artificial contrast variables and mutual information. J Chromatogr B Analyt Technol Biomed Life Sci. 2012;910:149–55.

Mai RY, Zeng J, Meng WD, Lu HZ, Liang R, Lin Y, Wu GB, Li LQ, Ma L, Ye JZ, Bai T. Artificial neural network model to predict post-hepatectomy early recurrence of hepatocellular carcinoma without macroscopic vascular invasion. BMC Cancer. 2021;21:283.

Barbie DA, Tamayo P, Boehm JS, Kim SY, Moody SE, Dunn IF, Schinzel AC, Sandy P, Meylan E, Scholl C, Fröhling S, Chan EM, Sos ML, Michel K, Mermel C, Silver SJ, Weir BA, Reiling JH, Sheng Q, Gupta PB, Wadlow RC, Le H, Hoersch S, Wittner BS, Ramaswamy S, Livingston DM, Sabatini DM, Meyerson M, Thomas RK, Lander ES, Mesirov JP, Root DE, Gilliland DG, Jacks T, Hahn WC. Systematic RNA interference reveals that oncogenic KRAS-driven cancers require TBK1. Nature. 2009;462:108–12.

Hänzelmann S, Castelo R, Guinney J. GSVA: gene set variation analysis for microarray and RNA-seq data. BMC Bioinformatics. 2013;14:7.

Chen X, Cubillos-Ruiz JR. Endoplasmic reticulum stress signals in the tumour and its microenvironment. Nat Rev Cancer. 2021;21:71–88.

Oakes SA, Papa FR. The role of endoplasmic reticulum stress in human pathology. Annu Rev Pathol. 2015;10:173–94.

Wang R. Hydrogen sulfide: the third gasotransmitter in biology and medicine. Antioxid Redox Signal. 2010;12:1061–4.

Yin P, Zhao C, Li Z, Mei C, Yao W, Liu Y, Li N, Qi J, Wang L, Shi Y, Qiu S, Fan J, Zha X. Sp1 is involved in regulation of cystathionine γ-lyase gene expression and biological function by PI3K/Akt pathway in human hepatocellular carcinoma cell lines. Cell Signal. 2012;24:1229–40.

Pan Y, Zhou C, Yuan D, Zhang J, Shao C. Radiation exposure promotes hepatocarcinoma cell invasion through epithelial mesenchymal transition mediated by H2S/CSE pathway. Radiat Res. 2016;185:96–105.

Bai KH, He SY, Shu LL, Wang WD, Lin SY, Zhang QY, Li L, Cheng L, Dai YJ. Identification of cancer stem cell characteristics in liver hepatocellular carcinoma by WGCNA analysis of transcriptome stemness index. Cancer Med. 2020;9:4290–8.

Liu CL, Pan HW, Torng PL, Fan MH, Mao TL. SRPX and HMCN1 regulate cancerassociated fibroblasts to promote the invasiveness of ovarian carcinoma. Oncol Rep. 2019;42:2706–15.

CAS   PubMed   Google Scholar  

Zabrenetzky V, Harris CC, Steeg PS, Roberts DD. Expression of the extracellular matrix molecule thrombospondin inversely correlates with malignant progression in melanoma, lung and breast carcinoma cell lines. Int J Cancer. 1994;59:191–5.

Isenberg JS, Roberts DD. THBS1 (thrombospondin-1). Atlas Genet Cytogenet Oncol Haematol. 2020;24:291–9.

PubMed   PubMed Central   Google Scholar  

Kaur S, Bronson SM, Pal-Nath D, Miller TW, Soto-Pantoja DR, Roberts DD. Functions of thrombospondin-1 in the tumor microenvironment. Int J Mol Sci. 2021;22:4570.

Guo D, Zhang D, Ren M, Lu G, Zhang X, He S, Li Y. THBS4 promotes HCC progression by regulating ITGB1 via FAK/PI3K/AKT pathway. FASEB J. 2020;34:10668–81.

Su F, Zhao J, Qin S, Wang R, Li Y, Wang Q, Tan Y, Jin H, Zhu F, Ou Y, Cheng Z, Su W, Zhao F, Yang Y, Zhou Z, Zheng J, Li Y, Li Z, Wu Q. Over-expression of Thrombospondin 4 correlates with loss of miR-142 and contributes to migration and vascular invasion of advanced hepatocellular carcinoma. Oncotarget. 2017;8:23277–88.

Sueyoshi T, Moore R, Sugatani J, Matsumura Y, Negishi M. PPP1R16A, the membrane subunit of protein phosphatase 1beta, signals nuclear translocation of the nuclear receptor constitutive active/androstane receptor. Mol Pharmacol. 2008;73:1113–21.

Colas E, Perez C, Cabrera S, Pedrola N, Monge M, Castellvi J, Eyzaguirre F, Gregorio J, Ruiz A, Llaurado M, Rigau M, Garcia M, Ertekin T, Montes M, Lopez-Lopez R, Carreras R, Xercavins J, Ortega A, Maes T, Rosell E, Doll A, Abal M, Reventos J, Gil-Moreno A. Molecular markers of endometrial carcinoma detected in uterine aspirates. Int J Cancer. 2011;129:2435–244.

Cui Z, Wang J, Chen G, Li D, Cheng B, Lai Y, Wu Z. The upregulation of CLGN in hepatocellular carcinoma is potentially regulated by hsa-mir-194-3p and associated with patient progression. Front Oncol. 2023;12:1081510.

Zhan K, Yang X, Li S, Bai Y. Correlation of endoplasmic reticulum stress patterns with the immune microenvironment in hepatocellular carcinoma: a prognostic signature analysis. Front Immunol. 2023;14:1270774.

Zhang Y, Zhang Z. The history and advances in cancer immunotherapy: understanding the characteristics of tumor-infiltrating immune cells and their therapeutic implications. Cell Mol Immunol. 2020;17:807–21.

Bagchi S, Yuan R, Engleman EG. Immune checkpoint inhibitors for the treatment of cancer: clinical impact and mechanisms of response and resistance. Annu Rev Pathol. 2021;16:223–49.

Ezhilarasan D. Hepatic stellate cells in the injured liver: perspectives beyond hepatic fibrosis. J Cell Physiol. 2022;237(1):436–49.

Chen Y, Qian B, Sun X, Kang Z, Huang Z, Ding Z, Dong L, Chen J, Zhang J, Zang Y. Sox9/INHBB axis-mediated crosstalk between the hepatoma and hepatic stellate cells promotes the metastasis of hepatocellular carcinoma. Cancer Lett. 2021;499:243–54.

Liberzon A, Birger C, Thorvaldsdóttir H, Ghandi M, Mesirov JP, Tamayo P. The Molecular signatures database (MSigDB) hallmark gene set collection. Cell Syst. 2015;1:417–25.

Lin Z, Huang W, He Q, Li D, Wang Z, Feng Y, Liu D, Zhang T, Wang Y, Xie M, Ji X, Sun M, Tian D, Xia L. FOXC1 promotes HCC proliferation and metastasis by upregulating DNMT3B to induce DNA hypermethylation of CTH promoter. J Exp Clin Cancer Res. 2021;40:50.

Download references

Acknowledgements

Not applicable.

This work was supported by grants provided by the National Natural Science Foundation of China (No. 81070125, 81270213, 81670306); the Science and Technology Foundation in Guangdong Province (No. 2010B031600032, 2014A020211002); the National Natural Science Foundation of Guangdong Province (No. 2017A030313503); the Science and Technology Foundation in Guangzhou City (No. 201806020084); the Fundamental Research Funds for the Central Universities (No. 13ykzd16, 17ykjc18); the Futian District Health and Public Welfare Research Project of Shenzhen City (No. FTWS2019001, FTWS2021016, FTWS2022018, FTWS2023064), the Shenzhen Fundamental Research Program (No. JCYJ20190808101405466, JCYJ20210324115003008, JCYJ20220530144404009).

Author information

Zhaorui Cheng, Shuangmei Li and Shujun Yang contributed equally to this work.

Authors and Affiliations

Department of Emergency, The Eighth Affiliated Hospital of Sun Yat- sen University, Shenzhen, 518003, Guangdong, P. R. China

Zhaorui Cheng, Shuangmei Li, Shujun Yang, Huibao Long, Haidong Wu, Xuxiang Chen & Tong Wang

The First Affiliated Hospital of Jiangxi Medical College, Nanchang University, Nanchang, 330000, Jiangxi, P. R. China

Xiaoping Cheng

You can also search for this author in PubMed   Google Scholar

Contributions

ZC, TW – designed study protocol, writing the manuscript; SL – construction and validation of ANN prediction model in HCC; SY – cell study in vitro; ZC – statistical data analysis; HL, HW, XC, XC – searching of world literature and discussion of the obtained results with the results of previous research; All the authors approved the final version of the article to be published.

Corresponding author

Correspondence to Tong Wang .

Ethics declarations

Ethics approval and consent to participate, consent for publication, competing interests.

The authors declare that there are no competing interests.

Additional information

Publisher’s note.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ . The Creative Commons Public Domain Dedication waiver ( http://creativecommons.org/publicdomain/zero/1.0/ ) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Cite this article.

Cheng, Z., Li, S., Yang, S. et al. Endoplasmic reticulum stress promotes hepatocellular carcinoma by modulating immunity: a study based on artificial neural networks and single-cell sequencing. J Transl Med 22 , 658 (2024). https://doi.org/10.1186/s12967-024-05460-9

Download citation

Received : 15 February 2024

Accepted : 01 July 2024

Published : 15 July 2024

DOI : https://doi.org/10.1186/s12967-024-05460-9

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Hepatocellular carcinoma
  • Endoplasmic reticulum stress
  • Artificial neural network
  • Immune infiltration

Journal of Translational Medicine

ISSN: 1479-5876

  • Submission enquiries: Access here and click Contact Us
  • General enquiries: [email protected]

research studies on stress

COMMENTS

  1. STRESS AND HEALTH: Psychological, Behavioral, and Biological

    LIFE STRESS, ANXIETY, AND DEPRESSION . It is well known that first depressive episodes often develop following the occurrence of a major negative life event (Paykel 2001).Furthermore, there is evidence that stressful life events are causal for the onset of depression (see Hammen 2005, Kendler et al. 1999).A study of 13,006 patients in Denmark, with first psychiatric admissions diagnosed with ...

  2. The impact of stress on body function: A review

    Some studies have shown that stress has many effects on the human nervous system and can cause structural changes in different parts of the brain (Lupien et al., 2009 [ 65 ]). Chronic stress can lead to atrophy of the brain mass and decrease its weight (Sarahian et al., 2014 [ 100 ]).

  3. Mind and Body Approaches for Stress and Anxiety: What the Science Says

    A 2017 meta-analysis looked at 24 studies—484 participants total—on heart rate variability (HRV) biofeedback and general stress and anxiety. The meta-analysis found that HRV biofeedback is helpful for reducing self-reported stress and anxiety, and the researchers saw it as a promising approach with further development of wearable devices ...

  4. Psychological and biological resilience modulates the effects of stress

    The participant cohort included 444 community adults between the ages of 18-50 in the greater New Haven, CT area who volunteered to participate in a study examining the role of stress and self ...

  5. Stress and Health: A Review of Psychobiological Processes

    The cumulative science linking stress to negative health outcomes is vast. Stress can affect health directly, through autonomic and neuroendocrine responses, but also indirectly, through changes in health behaviors. In this review, we present a brief overview of (a) why we should be interested in stress in the context of health; (b) the stress response and allostatic load; (c) some of the key ...

  6. Focus: The Science of Stress: Introduction: The Science of Stress

    The results of these studies emphasize a need to study intergenerational transmission of stress and its long-term effects. ... In this issue, the biological and social dynamics of stress are examined. Original research, reviews, and perspectives are presented on how stress affects development, metabolism, and various cellular and organ level ...

  7. Best practices for stress measurement: How to measure psychological

    Epidemiological studies confirm that both experiencing a greater number of stressful events and reporting high perceived stress over long periods of time are associated with worse mental and physical health, and mortality (Epel et al., 2018).The association between greater stressor exposure and increased disease risk has been replicated with many different types of stressor exposures (e.g ...

  8. Recent developments in stress and anxiety research

    Coinciding with WASAD's 3rd International Congress held in September 2021 in Vienna, Austria, this journal publishes a Special Issue encompassing state-of-the art research in the field of stress and anxiety. This special issue collects answers to a number of important questions that need to be addressed in current and future research.

  9. The science of stress relief

    Stress is a universal experience. In low amounts or short durations, stress can be helpful and could serve as a motivator or to increase our resilience. Studies have shown the health benefits of short-term exposure to acute physiological stressors. According to a study in BMC Medicine, a short duration of thermal stress, as might be achieved ...

  10. 6 things researchers want you to know about stress

    1. Stress can manifest in the body. While stress can seem like a mental and emotional experience, its effects on the body are well-documented. We've all felt how short-term stress, like being startled, can make the heart race. And ongoing stress can accumulate, causing inflammation, wearing on the immune system, and overexposing the body to ...

  11. Does the perception that stress affects health matter? The ...

    Objective: This study sought to examine the relationship among the amount of stress, the perception that stress affects health, and health and mortality outcomes in a nationally representative sample of U.S. adults. Methods: Data from the 1998 National Health Interview Survey were linked to prospective National Death Index mortality data through 2006.

  12. The Social Psychology of Stress, Health, and Coping

    Abstract. The study of stress and health is one of the richest areas of research in the social and biomedical sciences. In this chapter, we first describe core concepts in the study of stress, coping, and health. Second, we summarize key theoretical perspectives that frame social psychological research on stress and health.

  13. Full article: The impact of stress on students in secondary school and

    This study demonstrates the importance of protective social factors in mediating the effects of academic-related stress. In a cross-sectional study of tertiary nursing students from the United States, those with higher self-reported stress had higher incidence of substance use.

  14. STRESS RESEARCH

    80% of US workers experience work stress because of ineffective company communications. 39% of North American employees report their workload the main source of the work stress. 49% of 18 - 24 year olds who report high levels of stress felt comparing themselves to others is a stressor. 71% of US adults with private health insurance say the ...

  15. Stress effects on the body

    Stress affects all systems of the body including the musculoskeletal, respiratory, cardiovascular, endocrine, gastrointestinal, nervous, and reproductive systems. ... Explore how scientific research by psychologists can inform our professional lives, family and community relationships, emotional wellness, and more. ... Some studies show that an ...

  16. Full article: Academic stress as a predictor of mental health in

    A study by Bashir et al. (Citation 2019) examined academic stress in university students and found that high expectations from professors and competition among peers can increase stress in the university environment. The study points out that stress can have negative effects on overall well-being, affecting students' mental and physical health.

  17. How stress affects your health

    Reducing your stress levels can not only make you feel better right now, but may also protect your health long-term. Several research studies have demonstrated, for example, that interventions to improve psychological health can have a beneficial impact on cardiovascular health. As a result, researchers recommend boosting your positive affect ...

  18. Measurement of Human Stress: A Multidimensional Approach

    Stress is a multidimensional construct that comprises exposure to events, perceptions of stress, and physiological responses to stress. Research consistently demonstrates a strong association between stress and a myriad of physical and mental health concerns, resulting in a pervasive and interdisciplinary agreement on the importance of investigating the relationship between stress and health.

  19. A Stressful New Decade: The latest information on how stress shapes our

    A study published in 2018 suggests excess noradrenaline "can contribute to arterial hypertension and predisposes to cardiovascular and renal damage." More recently, a groundbreaking study suggested that stress is a direct cause of hair-graying, confirming an age-old suspicion. The culprit was found to be one of the three hormones we just ...

  20. Enhancing psychological well-being in college students: the mediating

    The prevalence of depression among college students is higher than that of the general population. Although a growing body of research suggests that depression in college students and their potential risk factors, few studies have focused on the correlation between depression and risk factors. This study aims to explore the mediating role of perceived social support and resilience in the ...

  21. Frontiers

    Therefore, this study aimed to Blanco et al. (2021) explore the associations of optimism with sleep quality, stress, and mental health; and Trautmann et al. (2016) ascertain whether sleep and stress mediate the connection between optimism and mental health, providing valuable insights into a promising intervention strategy regarding elevating ...

  22. Stress-related cell damage linked to negative mental and physical

    Mar. 22, 2023 — New research finds that simply anticipating stress related to political elections causes adverse physical health effects. However, the study also finds there is something people ...

  23. Research Shows That 1 in 5 U.S. Farmers Binge Drink During Stress

    A new University of Georgia study reveals that one in five U.S. farmers report binge drinking when they experience high levels of stress. The post Research shows that 1 in 5 U.S. farmers binge ...

  24. Clinical Psychological Assessment of Stress: A Narrative Review of the

    The Job Stress Survey investigates generic sources of stress that take into account the frequency and severity of perceived stress, in line with the studies (Peter & Siegrist, 1999; Siegrist et al., 2004) that highlight how the temporal aspect can be a significant variable for the professional environment. According to the literature, the JSS ...

  25. Post-Traumatic Stress Disorder (PTSD)

    Prevalence of Post-Traumatic Stress Disorder Among Adolescents. Based on diagnostic interview data from National Comorbidity Survey Adolescent Supplement (NCS-A), Figure 3 shows lifetime prevalence of PTSD among U.S. adolescents aged 13-18. 4 An estimated 5.0% of adolescents had PTSD, and an estimated 1.5% had severe impairment.

  26. Study: Link Between Ag Stress, Alcohol

    LINCOLN, Neb. (DTN) -- About one in four farmers in the United States turn to binge drinking in response to high stress levels, according to a new study conducted by researchers at the University ...

  27. July: dogs-stress-study

    Dogs experience emotional contagion from the smell of human stress, leading them to make more 'pessimistic' choices, new research finds. The University of Bristol-led study, published in Scientific Reports today [22 July], is the first to test how human stress odours affect dogs' learning and emotional state.

  28. Best practices for stress measurement: How to measure psychological

    Introduction. Epidemiological studies confirm that both experiencing a greater number of stressful events and reporting high perceived stress over long periods of time are associated with worse mental and physical health, and mortality (Epel et al., 2018).The association between greater stressor exposure and increased disease risk has been replicated with many different types of stressor ...

  29. Atlanta airport ranks #2 for stress-free travel, study says

    The study analyzed 30 major airports using 2023 data from the Bureau of Transportation. The ranking considered factors like average fares, delayed, diverted, and canceled flights, TSA wait times ...

  30. Endoplasmic reticulum stress promotes hepatocellular carcinoma by

    Hepatocellular carcinoma (HCC) is characterized by the complex pathogenesis, limited therapeutic methods, and poor prognosis. Endoplasmic reticulum stress (ERS) plays an important role in the development of HCC, therefore, we still need further study of molecular mechanism of HCC and ERS for early diagnosis and promising treatment targets. The GEO datasets (GSE25097, GSE62232, and GSE65372 ...