• Open access
  • Published: 05 February 2022

Work environment risk factors causing day-to-day stress in occupational settings: a systematic review

  • Junoš Lukan 1 , 2   na1 ,
  • Larissa Bolliger 3   na1 ,
  • Nele S. Pauwels 4 , 5 ,
  • Mitja Luštrek 1 , 2 ,
  • Dirk De Bacquer 3 &
  • Els Clays 3  

BMC Public Health volume  22 , Article number:  240 ( 2022 ) Cite this article

10k Accesses

17 Citations

13 Altmetric

Metrics details

While chronic workplace stress is known to be associated with health-related outcomes like mental and cardiovascular diseases, research about day-to-day occupational stress is limited. This systematic review includes studies assessing stress exposures as work environment risk factors and stress outcomes, measured via self-perceived questionnaires and physiological stress detection. These measures needed to be assessed repeatedly or continuously via Ecological Momentary Assessment (EMA) or similar methods carried out in real-world work environments, to be included in this review. The objective was to identify work environment risk factors causing day-to-day stress.

The search strategies were applied in seven databases resulting in 11833 records after deduplication, of which 41 studies were included in a qualitative synthesis. Associations were evaluated by correlational analyses.

The most commonly measured work environment risk factor was work intensity, while stress was most often framed as an affective response. Measures from these two dimensions were also most frequently correlated with each other and most of their correlation coefficients were statistically significant, making work intensity a major risk factor for day-to-day workplace stress.

Conclusions

This review reveals a diversity in methodological approaches in data collection and data analysis. More studies combining self-perceived stress exposures and outcomes with physiological measures are warranted.

Peer Review reports

Over the past decades, substantial attention has been directed to research focusing on chronic exposure to stressors in occupational settings and the adverse impact of stress on chronic disease outcomes [ 1 , 2 ]. Psychosocial risk factors have been the last ones to be considered, but are now widely accepted to be as important as other factors like biological or chemical risks [ 3 ]. The influence on mental and cardiovascular health in particular has been confirmed and explained through frameworks at the forefront of stress research, such as the Job Demand-Control-Support model [ 4 – 6 ], the Effort-Reward Imbalance model [ 7 ], and the Job Demands-Resources model [ 8 ].

Evidence of chronic stressors influencing workers’ health and well-being is accumulating, and several systematic reviews and meta-analyses with a focus on studies investigating such relations are available. There is evidence that psychosocial stress is associated with cardiovascular morbidity and mortality [ 1 ], musculoskeletal disorders [ 9 ], mental health problems, such as depression and anxiety [ 2 ], and health risk behaviour, such as cigarette smoking, alcohol consumption, and overweight [ 10 ]. The most commonly studied workplace risk factors include job demand and job control (such as in the Job Demand-Control-Support model) [ 1 ], but several others have been studied, such as job insecurity, procedural (in)justice, workplace conflict or bullying [ 2 ], and workplace violence [ 11 ]. A lot of these risk factors are structural and as such measured at a single time point. On the other hand, a preliminary literature search revealed little information about how these structural risk factors manifest in daily work life and what the specific (if any) work environment risk factors causing day-to-day (i.e., non-chronic) stress are.

Understanding how day-to-day work situations lead to the experience of stress is important for several reasons. Stress measurement often relies on self-reports, which are subject to memory bias [ 12 ] and there are indications that chronic assessments are not simply the sum of multiple moment-to-moment ratings [ 13 ]. When stress is measured several times a day, these repeated measurements can instead capture stressful situations during or soon after they occur and the risk of memory bias is reduced. This relationship would also be important to understand since, in order to test hypotheses about how particular stressors lead to health outcomes, temporal relationships need to be explored [ 14 , 15 ]. Finally, the question of how to design stress management interventions is still open [ 16 ]. Broad constructs such as work demands and decision latitude are relatively stable aspects of a job and translating the findings from chronic stress research into stress management strategies applicable in every-day working life would require a better understanding of their day-to-day manifestations. To the best of our knowledge, no systematic review with a focus on day-to-day stress and related work environment risk factors has been performed so far.

We were interested in how various situations at work translate to an experience of stress and which situations are the most important for this experience. We named this relationship ‘day-to-day stress’, which differs from chronic stress in that it can entail daily situations in addition to structural characteristics of the workplace. Day-to-day stressors do not necessarily have long-lasting consequences but do influence the perception of work environment and elicit some kind of response from a person. We do not presume any conceptual difference between day-to-day stress and chronic stress or stressors, but rather differentiate them based on the methodology of stress measurement. As such, we did not restrain our selection of studies to review to any particular definition of (day-to-day or chronic) stress. To understand how exposures to work environment risk factors are related to daily variations of stress, we instead focused on the studies that measure these repeatedly with self-reports or continuously using physiological measurements and in real-world occupational settings.

The objective of this systematic review was to explore the onset of day-to-day stress by summarising evidence on potential day-to-day work environment risk factors (stressors), which have an immediate effect on self-perceived stress levels or physiological stress responses, and which may or may not cause chronic stress.

Materials and methods

We conducted this systematic review by following the Cochrane Handbook for Systematic Reviews [ 17 ] and the PRISMA 2009 Checklist [ 18 ] and registered it on PROSPERO under the ID: CRD42018105355 [ 19 ].

Eligibility criteria

We first settled on working definitions of the concepts of interest. Throughout this systematic review, we use the term work environment risk factors for stressors, which we defined more specifically as causes or predictors of stress, potentially occurring on a day-to-day basis within occupational settings . Furthermore, we defined self-perceived stress outcomes as consequences of such work environment risk factors which were measured with self-perception-based scales, questionnaires, or surveys .

As the main eligibility criterion, the studies needed to include work environment risk factors and either self-perceived stress outcomes or physiological stress detection or both. We looked for any (objective) descriptions of work situations and their consequences in terms of stress. We did not constrain the outcomes to any particular definition of stress as long as the authors of the original studies framed them as somehow related to the phenomenon of psychosocial stress.

Both, risk factors and outcomes, were required to be measured repeatedly, so methods capable of producing repeated or continuous measurements had to be used. One such method is Ecological Momentary Assessment [ 20, EMA ], which is a research method allowing participants to report their experiences in real-time and in real-world settings, in which data are collected repeatedly (i.e., more than two measurement points) over time and often through a digital platform such as a smartphone application [ 21 ]. Moreover, the phenomenon of interest was day-to-day stress, so studies focusing on chronic stress only were not considered.

We focused on studies set in a real-world working environment, and either the workplace setting or the occupational profile needed to be extractable. Healthy full-time and part-time workers of working age were chosen as the population of interest. Observational quantitative and mixed-methods studies (where only the quantitative part of the latter was of relevance) including at least two measurement points were included.

Strategies for database searching

We devised a search strategy according to the eligibility criteria described in the previous subsection. The main blocks of the search strategy require that 1) a study deals with stress (or synonymous concepts), which should furthermore be 2) day-to-day or episodic (or similar). We also set 3) the requirements for methods, which could be either ecological momentary assessments or other methods capable of producing repeated measures, and 4) that the setting of interest is the work setting. The full search strategies with indexing terms and free text words for all the databases can be found in Supplementary Figs. 1 to 7 [see Additional file  3 ].

We evaluated the search strategy with the PRESS checklist [ 22 ] before we applied it in the following databases: PubMed, Embase, Web of Science, Scopus, CINAHL, ERIC, and PsycArticles. We first carried out a search on 31 August 2018 and later did an update of the initial search on 3 July 2020. The only limitation we used at the time of the searches was “English language”.

Study selection process

The studies were selected based on the process described in the PRISMA statement [ 18 ]. After merging all the results and manually deduplicating them, we screened the titles and abstracts, and evaluated the full text of the remaining articles for eligibility.

Both title and abstract screening and full-text evaluation were done independently by two authors using the Rayyan software [ 23 ]. Conflicts after both screening phases were discussed until consensus was reached. We followed the same procedure when we updated our search.

Quality assessment strategy

The quality assessment in systematic reviews includes two phases: 1) evaluation of the quality at study level (each study separately) and 2) evaluation of the body of evidence (all included studies together) to give a thorough quality estimate of the evidence at hand.

For the quality assessment at study level, we used the QualSyst tool for quantitative research [ 24 ]. This tool offers ‘N/A’ grades for criteria that are inapplicable. Out of the 14 criteria in the tool, we omitted three—random allocation to treatment group (1) and blinding of investigators (2) and subjects (3)—since they are only applicable to intervention studies.

Consequently, each study was assessed according to 11 criteria (e.g., question or objective sufficiently described) on a three point scale. From these, the summary score is calculated, which is a number between 0 and 1, where 1 denotes complete satisfaction of all applicable criteria. This procedure was done independently by two authors and any conflicts were discussed until consensus was reached.

The quality assessment of the body of evidence was evaluated using the GRADE approach [ 25 ], including five criteria for downgrading (risk of bias, inconsistency, indirectness, imprecision, and publication bias) and two criteria for upgrading (large magnitude of effect and dose-response gradient) before an overall score was given.

Study selection

We applied the search strategy as discussed in Strategies for database searching and retrieved 15362 records. We removed duplicates first and then eliminated irrelevant studies by first screening their titles and abstracts and then considering the full text of a selected subset.

We followed the same procedure for the second search, in which we retrieved 18996 records. We excluded all the records already found in the first search as well as some duplicates from different databases, then repeated the same screening procedure. The whole process is illustrated in Fig.  1 .

figure 1

Study selection presented with the PRISMA flow diagram [ 18 ]. The n 1 and n 2 refer to the first search we performed on 31 August 2018 and to the update search from 3 July 2020, respectively. Their sum is reported as n

As per recommendations in Moher et al. [ 18 ], we noted the exclusion reasons for all the publications we reviewed in full. 149 studies were excluded based on the main eligibility criterion: either the work environment risk factors or stress outcomes were missing, or they were not measured repeatedly. In 35 cases, neither the workplace setting nor the occupational profile was mentioned. In 26 cases, no work environment risk factor could be extracted. 19 publications turned out to be abstracts only. Four further studies were excluded because not all participants were described as healthy. Three studies were not accessible: two PhD theses [ 26 , 27 ], of which authors were contacted, but we got no response, and a report [ 28 ], the author of which is deceased. Two additional duplicates were found, one study only included observational methods (i.e., no self-reports), and one study was available in Korean only.

After this, we were left with 41 studies (34 from the first and 7 from the second search), which were included in the final review for a qualitative synthesis.

Study characteristics

Table  1 lists all the studies included in the final review and gives their most important characteristics. In addition to the number of subjects included ( N ) and the work setting of the study, it includes the study duration and assessment frequency. Added are the quality rating of the study and information about whether data analysis included multilevel analysis.

The studies included from N =14 to N =304 participants, with the median N median =83. The study duration also varied, usually together with the assessment frequency. Thus, some studies only looked at one day, but measured some parameters continuously (e.g., blood pressure in study ID 32), whereas others only sampled once a day for 60 days (study ID 26) or even once a week, but lasted for 182 days (study ID 37).

The average quality of the studies according to QualSyst was M QualSyst =0.86. Specifically, 10 out of the 41 studies were rated at the highest quality level, 28 studies were rated between 0.99 and 0.51, and 3 studies were rated as 0.50 or below. Points were most often deducted because recruitment methods were not clearly reported (‘Method of subject selection’ criterion) or when the authors only reported significance of the main results, but no estimate of variance, such as confidence intervals or standard errors (‘Some estimate of variance’ criterion).

The GRADE approach rates randomised trials as ‘high quality’, observational studies as ‘low quality’, and any other evidence as ‘very low quality’. Since all included studies of this review are observational studies, the quality of the body of evidence was consequently rated as low. Additionally, according to the criteria for downgrading or upgrading mentioned in Quality assessment strategy, no downgrading was required and no upgrading was feasible for our body of evidence, so ‘low’ was also the final score.

Work environment risk factors, self-perceived stress outcomes, and physiological measurements

We identified work environment risk factors, self-perceived stress outcomes, and physiological measurements measured in each study. We extracted all constructs that were measured more than twice (i.e., repeatedly or continuously), while one-time measurements (e.g., during initial baseline screening) were not considered. Since concepts like stress are defined in different ways across the studies, we also looked at the measurement instruments. The work environment risk factors and stress outcomes extracted in this way are published as Supplementary Tables 1 and 2 [see Additional file  1 ].

We classified these into broader categories, which were based on established frameworks. For work environment risk factors, we used the 6th European Working Conditions Survey classification of job quality indices [ 29 ] as the basis. These are objective features of a job which have been proven to have an impact on the health and well-being of workers. They are also only weakly correlated, so that they function as independent descriptions of job quality. Hence, these served well to get a higher level overview of work environment risk factors.

Similarly, stress outcomes were classified according to the stress model described by Ice and James [ 30 ]. They describe stress as a combination of affective, behavioural, and physiological responses, which impact mental health and can have physical health outcomes. These responses are consequences of a person’s appraisal of stressors (stimuli). While classifying the studies in our review, we also identified the need for two additional consequences of stress. First, stress can also affect a person’s cognition, not only at the stage of appraisal, but as its consequence; forgetting of intentions and cognitive failure are examples of this outcome. Second, we determined motivational responses, specifically work engagement, to be sufficiently distinct from affective responses to deserve its own category. This outcome involves not only emotions, but goal-directed behaviour closely related to affective responses.

The frequencies of different measures of work environment risk factors and stress outcomes are summarised in Table  2 . The risk factor most often measured in the included studies was work intensity, defined for example as time pressure or job demand. This was followed by social environment risk factors (such as co-worker and supervisor support) and ‘various’ factors (such as the number or type of stressful situations). On the other end, affective responses were by far the most often measured outcomes, especially as assessed by the Positive and Negative Affect Scale [ 31, PANAS ]. Some studies looked at physiological responses to stress as well, while other outcomes were rarely considered.

It is important to note that while all studies included at least one work environment risk factor and one stress outcome as a consequence of the design of our eligibility criteria, each study could measure more than one risk factor or outcome. This means that the total number of measurements is larger than the number of included studies.

Correlation coefficients

The variables of work environment risk factors, self-perceived stress outcomes, and physiological measurements were considered in the relation structure widely known as ‘exposure/predictor–outcome’ or ‘independent variable–dependent variable’.

As shown in Table  1 , 28 studies analysed their data by using multilevel models, while others resorted to other analyses, such as t -tests, correlational tables, and descriptive analysis. Since multilevel models control for dependencies between predictors (usually work environment risk factors) they give a more complete insight into relationships between all modelled variables. But for these studies direct comparison of coefficient estimates would not be appropriate since the models’ structure varied from study to study.

To produce a meaningful comparison, we therefore decided to focus on correlation coefficients in our synthesis. As these were available in the 28 studies as well as some others, they enabled us to compare more directly the results of more studies. It needs to be noted, however, that only bivariate relationships are reflected in these analyses, while more complex relationships between variables are omitted.

Accordingly, correlation coefficients between work environment risk factors (exposures) and self-perceived stress outcomes (outcomes), and work environment risk factors (exposures) and physiological measurements (outcomes) are available in Supplementary Table B [see Additional file  2 ].

For studies that did multilevel analysis, both within-person level and between-persons level results were extracted. But for other studies only a part of the results of interest were extractable. For example, only the results of between-persons level were reported in the study ID 37 or correlation coefficients were not reported for all variables included in the study ID 33 (e.g., no correlation coefficients between nursing tasks and heart rate).

Figure  2 shows the number of statistically significant and nonsignificant correlation coefficients for each pair of work environment risk factor and stress outcome. In this figure, we focused on within-subject correlations only as these were the ones that can capture day-to-day variation of stressors and responses.

figure 2

The frequency of significant and nonsignificant correlation coefficients between categories of work environment risk factors and stress outcomes

As mentioned in Work environment risk factors, self-perceived stress outcomes, and physiological measurements (see Table  2 ), the most commonly measured risk factor and outcome were work intensity and affective response, respectively. Correspondingly, their relationship in the form of correlation was also the most commonly reported one. Note that the number of correlation coefficients does not directly follow from the number of studies studying a certain relationship, since Table  2 shows measures of all studies, regardless of whether they did multilevel analysis and whether within-subject correlations could be extracted. Additionally, a study looking at several measures of work environment risk factors or stress outcomes could report a correlation within each pair, so that the number of correlations can be even higher than the number of different measures.

Furthermore, most of the correlations between work intensity and affective response were statistically significant. Affective response was also commonly correlated with social environment and this relationship was more often statistically significant than not.

On the other hand commuting from and to the workplace was not significantly correlated with affective responses most of the time. Interestingly, the second most common type of outcomes, physiological responses, were mostly not significantly correlated with any of the risk factors.

In general, the correlations reported in the included studies were more often significant than not (110 significant vs. 80 nonsignificant correlations), but they were typically low. Only two of all the within-subject correlations exceeded 0.5 in their absolute magnitude.

Among the studies that included within-subject correlation coefficients in their results, the most commonly measured work environment risk factor was work intensity, and it was correlated most often with affective response. The high frequency can be to some extent explained by our search strategy. Both stress and demand were included as search terms and we categorised stress(fulness) as an affective response and work demand as work intensity. But since definitions of stress are diverse and often include concepts such as negative affect, our review captured many other stress responses as well as work environment risk factors (stressors).

Work intensity is often measured in epidemiological studies about health consequences of chronic stress, such as coronary heart disease [ 32 ] and depression [ 33 ]. This is reflected in the high number of studies that included it as the stressor of interest. However, since work intensity is most often paired with the control dimension to describe job strain, such as in the Job Demand-Control-Support model, the relative rarity of correlations between skills and discretion and stress outcomes is surprising. It seems that the control dimension of this model is relatively less well researched. However, it is still unclear whether demands and control are related to stress and its health consequences independently, or whether their interaction in the form of job strain is more important [ 34 ]. To settle this question, it would be crucial to explore the role of the control dimension more carefully.

The social environment (e.g., co-worker and supervisor support) has also been correlated with affective and other responses in the included studies. This should be beneficial for gathering more evidence for the relationship of the support dimension with health outcomes, for which only limited evidence has been available in existing reviews [ 33 ]. A similar statement could be made for the prospects category, which has found its place in the Effort-Reward Imbalance model [ 7 ].

Another surprising result was that physiological responses were generally not statistically significantly correlated with any of the work environment risk factors, despite the fact that stress is often predicted from physiological parameters [ 35 ]. This can be explained either with the studies’ analyses choices or complexity of their models. Some of the studies that measured physiological responses did not perform multilevel analysis, but simpler statistical analyses such as t -tests and analysis of variance (studies ID 17, 31, and 32). Others did perform regression analysis or multilevel analysis, but did not report within-subject correlation coefficients for physiological parameters (studies ID 14, 33, and 38). Only three studies reported within-subject correlation coefficients and were included in Fig.  2 , but the relationships between work environment risk factors were usually too complex to be captured by these simple coefficients. For example, co-worker support mediated the daily trajectory of some parameters of heart rate variability [ 36, study ID 8 ] and the relationship between work-to-family conflict and cortisol slope from dinner to bedtime was mediated by supervisor support [ 37, study ID 34 ].

This relative rarity of studies dealing with physiological aspects of stress compared to the field of chronic stress research can be put into a broader context with the help of the conceptual framework proposed by Martikainen et al. [ 38 ] and adapted by Rugulies [ 39 ]. This framework describes the connection between different levels of work environment and health outcomes. It starts with the broadest, (i) macro-level, economic, social, and political structures, and continues through (ii) meso-level workplace structures, (iii) meso-level psychosocial working conditions, to (iv) individual-level experience and cognitive and emotional processes. The latter elicit either (v) psycho-physiological changes or (vi) health-related behaviours, which in turn impact (vii) workers’ health and illness.

It is well established that meso-level workplace structures (ii), such as job insecurity [ 40 ], and meso-level psychosocial working conditions (iii), such as job strain [ 41 ], are related to the risk of diseases and disorders (vii). This has been observed both in immediate physiological responses to stress as well as sustained physiological and behaviour changes. For example, longer duration of work-related stress results in increased rise in morning cortisol level and reduced heart rate variability, and acute stress response involves elevated blood pressure [ 42 ]. On the other hand, job strain has been found to be linked to hypertension, atherosclerosis, and smoking intensity [ 43 ].

Some of the mechanisms of how this happens are also understood. First, pathophysiological effects of stress (v) have been detailed [ 44 ], such as neuroendocrine mechanisms of elevated cortisol and catecholamine (epinephrine) levels as well as inhibited anabolism. Second, stress is related to altered behaviour (vi), such as smoking and alcohol consumption [ 10 ], where this is seen as a second ‘indirect’ pathway of the link between stress and stress-related diseases.

A causal relationship between stress and cardiovascular diseases has still not been established, however, and the pathological mechanisms of chronic and acute stress may differ [ 41 ]. This evidence gap might be owed to a poor understanding of how psychological processes (iv) are involved in this pathway. Steptoe and Kivimäki [ 41 ] explicitly limit the focus of their review to ‘exposure to external stressors, rather than on psychological and biological factors affecting vulnerability to adversity’ (p. 360). And while they mention that ‘one reason for the weak relationship between physiological stress responses and future disease is that mental stress testing measures a propensity to high- or low-stress responsivity’ [ 42, p. 341 ], they only admit this role to biological stress reactivity.

But it might be precisely the psychosocial factors, which Martikainen et al. [ 38 ] see as ‘mediating the effects of social structural factors on individual health outcomes’ (p. 1091), that is, the pathway from (ii) and (iii) through (iv) to (vii), where the key to better understanding this causal relationship lies. It might be through perceptions and psychological processes at the individual level that these macro- and meso-level social processes lead to direct psychobiological processes or modified health-related behaviours and lifestyles and, in turn, influence health [ 38 ].

Some of the studies included in our review deal with the relationships in this pathway and more systematic research is needed. This also has implications for planning interventions better, since effects of stress and depression management techniques on cardiac outcomes are still uncertain [ 41 ]. While it is clear that occupational stress increases risk for coronary heart disease [ 45 ], more research is needed on how to lower this risk. For example, it is possible to modify work schedule to ameliorate exposure to job strain, but only a randomized clinical trial which would test this intervention could truly assess its effect [ 43 ].

Strengths and limitations

As illustrated, different methods of data collection (e.g., time span, number of measurement points) and a wide range of data analysis approaches (e.g., descriptive results, multilevel analysis) were used across the included studies. This heterogeneity led to a challenging data synthesis and study comparison, and restricted us to a qualitative (narrative) synthesis, rather than a quantitative synthesis in the form of a meta-analysis.

To enable a meaningful comparison of the studies’ results, we needed to introduce a rigorous approach to summarising them. First, we developed a working definition of the concepts explored in our research question. Despite widely acknowledged psychosocial stress models [ 4 , 46 ], definitions of psychosocial stress and job stressors are not used consistently in the literature. This diversity in terminology of ‘stressor’ and ‘stress’ was apparent during several steps, such as during the construction of the search strategies and during data extraction. By framing these phenomena—for the purpose of this systematic review—as ‘work environment risk factors’ and ‘self-perceived stress outcomes’, we attempted to harmonize these differences in terminology. To be able to study relationships between stressors and stress, we classified measured variables into one of these two categories according to our working definitions (see Background). This allowed us to compare different studies’ results.

As mentioned, we disregarded the original studies’ framing of independent and dependent variables and instead classified them according to our own criteria. While this enabled us to look at the studies from the point of view of our study question, it introduced the risk of misrepresenting the original findings.

This concern was alleviated by limiting our data extraction of study results to correlation coefficients, which helped us increase study comparability at the same time. The advantage of considering only correlational analyses is that the Pearson correlation coefficient is a symmetric statistic. This allowed us to sidestep the original authors’ hypotheses and models and frame their (partial) results in our work environment risk factors and stress outcomes research question. This had the effect of including studies from different fields and getting an overview of these relationships, which was as broad as possible.

On the other hand, considering only correlational analysis led to an incomplete representation of results of several included studies. This is especially true of the studies performing more extensive analyses, since correlational analysis was merely an intermediate step before final conclusions based on multilevel analysis or analyses focusing on moderating or mediating effects. This has already been illustrated with the case of physiological responses. With such a diverse set of predictors and outcomes, however, comparing the results of these more complex models proved to be problematic, since it is impossible to compare specific effects without considering the full model.

Another consideration is the focus on statistical significance of correlations. While a fixation on statistical significance has been widely criticised [ 47 ], it served as a good first step in comparisons of heterogeneous studies. Effect size examination made little sense as all the reported within-subject correlation coefficients were low (i.e., r <0.5). Noting the above points about simplification with regard to results reporting, the raw number of statistically significant and nonsignificant correlations should still serve as a first overview of the field.

While the field of chronic stress in the workplace is very well established, how daily work situations translate to day-to-day experience of stress and later to chronic conditions seems to be less understood.

We identified several high-quality studies dealing with this topic. The models they employ and the analytical methods they use are well developed. However, their research questions are particular and usually involve a somewhat narrow definition of stress. Instead of approaching stress outcomes as manifestations of a multifaceted response, only some types of responses are considered, most often affective responses. In our review, none of the studies approached this topic from a full-fledged stress model that would incorporate all the relevant aspects of the response to stressors.

Such a study would first require a combination of various data collection methods, such as ecological momentary assessment and continuous physiological monitoring. It would also call for a more complex analysis approach, such as combining multilevel modelling with structural equation modelling or other probabilistic graphical models. Finally, it would need to deal with the problem in the context of a well-established model of stress that lends itself well to such modelling.

Availability of data and materials

All data generated or analysed during this study are included in this published article and its Additional Files. Additional evaluations done by two reviewers independently are available from the corresponding author on reasonable request.

Abbreviations

Ecological momentary assessment

Positive affect negative affect scale

Fishta A, Backé E-M. Psychosocial stress at work and cardiovascular diseases: An overview of systematic reviews. Int Arch Occup Environ Health. 2015; 88(8):997–1014. doi:10.1007/s00420-015-1019-0.

Article   PubMed   PubMed Central   Google Scholar  

Harvey SB, Modini M, Joyce S, Milligan-Saville JS, Tan L, Mykletun A, Bryant RA, Christensen H, Mitchell PB. Can work make you mentally ill? a systematic meta-review of work-related risk factors for common mental health problems. Occup Environ Med. 2017; 74(4):301–10. doi:10.1136/oemed-2016-104015.

Article   Google Scholar  

Magnavita N, Chirico F. New and emerging risk factors in occupational health. Appl Sci. 2020; 10(24):8906. doi:10.3390/app10248906.

Article   CAS   Google Scholar  

Johnson JV, Hall EM. Job strain, work place social support, and cardiovascular disease: a cross-sectional study of a random sample of the Swedish working population. Am J Public Health. 1988; 78(10):1336–42. doi:10.2105/ajph.78.10.1336.

Article   CAS   PubMed   PubMed Central   Google Scholar  

Johnson JV, Hall EM, Theorell T. Combined effects of job strain and social isolation on cardiovascular disease morbidity and mortality in a random sample of the Swedish male working population. Scand J Work Environ Health. 1989; 15(4):271–79. doi:10.5271/sjweh.1852.

Article   CAS   PubMed   Google Scholar  

Karasek R, Theorell T. Healthy Work: Stress, Productivity and the Reconstruction of Working Life. New York: Basic books; 1990.

Google Scholar  

Siegrist J. Adverse health effects of high-effort/low-reward conditions. J Occup Health Psychol. 1996; 1(1):27–41. doi:10.1037/1076-8998.1.1.27.

Demerouti E, Bakker AB, Nachreiner F, Schaufeli WB. The job demands-resources model of burnout. J Appl Psychol. 2001; 86(3):499–512. doi:10.1037/0021-9010.86.3.499.

Hauke A, Flintrop J, Brun E, Rugulies R. The impact of work-related psychosocial stressors on the onset of musculoskeletal disorders in specific body regions: A review and meta-analysis of 54 longitudinal studies. Work Stress. 2011; 25(3):243–56. doi:10.1080/02678373.2011.614069.

Siegrist J, Rödel A. Work stress and health risk behavior. Scand J Work Environ Health. 2006; 32(6):473–81. doi:10.5271/sjweh.1052.

Article   PubMed   Google Scholar  

Magnavita N, Stasio ED, Capitanelli I, Lops EA, Chirico F, Garbarino S. Sleep problems and workplace violence: A systematic review and meta-analysis. Front Neurosci. 2019; 13(997):1–18. doi:10.3389/fnins.2019.00997.

Sato H, Kawahara J-i. Selective bias in retrospective self-reports of negative mood states. Anxiety Stress Coping. 2011; 24(4):359–67. doi:10.1080/10615806.2010.543132.

Bell C, Johnston D, Allan J, Johnston M, Pollard B. Repeated real time measures of work stress in nurses may not relate to questionnaire accounts. Psychol Health. 2013; 28(sup1 (EHPS 2013 Abstracts)):65–66. doi:10.1080/08870446.2013.810851.

Shiffman S, Stone AA. Introduction to the special section: Ecological momentary assessment in health psychology. Health Psychol. 1998; 17(1):3–5. doi:10.1037/h0092706.

Smyth JM, Stone AA. Ecological momentary assessment research in behavioral medicine. J Happiness Stud. 2003; 4(1):35–52. doi:10.1023/a:1023657221954.

Hurrell JJ. Organizational stress intervention In: Barling J., Kelloway E. K., Frone M. R., editors. Handbook of Work Stress. SAGE Publications, Inc.: 2005. p. 623–46. Chap. 27. http://dx.doi.org/10.4135/9781412975995.n27 .

Cochrane. Cochrane Handbook for Systematic Reviews of Interventions, 6.1 edn.: Cochrane; 2020.

Moher D, Liberati A, Tetzlaff J, Altman DG. Preferred reporting items for systematic reviews and meta-analyses: The PRISMA statement. PLoS Med. 2009; 6(7):1000097. doi:10.1371/journal.pmed.1000097.

Bolliger L, Lukan J, Pauwels NS, Luštrek M, De Bacquer D, Clays E. Work environment risk factors causing day-to-day stress in occupational settings: A systematic review. PROSPERO 2018 CRD42018105355. 2018. https://www.crd.york.ac.uk/prospero/display_record.php?ID=CRD42018105355 .

Csikszentmihalyi M, Larson R, Prescott S. The ecology of adolescent activity and experience. J Youth Adolesc. 1977; 6(3):281–94. doi:10.1007/bf02138940.

Gibbons CJ. Turning the page on pen-and-paper questionnaires: Combining ecological momentary assessment and computer adaptive testing to transform psychological assessment in the 21st century. Front Psychol. 2017; 7. http://dx.doi.org/10.3389/fpsyg.2016.01933.

McGowan J, Sampson M, Salzwedel DM, Cogo E, Foerster V, Lefebvre C. PRESS – Peer Review of Electronic Search Strategies: 2015 Guideline Explanation and Elaboration (PRESS E&E). Ottawa: Elsevier BV; 2016.

Ouzzani M, Hammady H, Fedorowicz Z, Elmagarmid A. Rayyan—a web and mobile app for systematic reviews. Syst Rev. 2016; 5(1). http://dx.doi.org/10.1186/s13643-016-0384-4.

Kmet L, Lee R, Cook L. Standard Quality Assessment Criteria for Evaluating Primary Research Papers from a Variety of Fields. Edmonton, Alta: Institute of Health Economics; 2004.

Schünemann H, BroŻek J, Guyatt G, Oxman A. Handbook for Grading the Quality of Evidence and the Strength of Recommendations Using the GRADE Approach. London: Cochrane; 2013. Cochrane. Retrieved from https://training.cochrane.org/resource/grade-handbook .

Heeren JK. Stress, hardiness, and burnout among psychiatric nurses working in hospital settings. Phd thesis: University of Virginia; 1991.

Northrop LME. Stress, social support, and burnout in nursing home staff. Phd thesis: West Virginia University; 1996.

Johnstone M. Time and tasks: Teacher workload and stress. Research Report ED368700. 1993. Spotlights 44.

Eurofound. Sixth European Working Conditions Survey – Overview Report (2017 Update). Luxembourg: Publications Office of the European Union; 2017. doi:10.2806/422172.

Ice GH, James GD. Conducting a field study of stress In: Ice G. H., James G. D., editors. Measuring Stress in Humans. Cambridge: Cambridge University Press: 2006. p. 3–24. Chap. 1.

Chapter   Google Scholar  

Watson D, Clark LA, Tellegen A. Development and validation of brief measures of positive and negative affect. J Pers Soc Psychol. 1988; 54(6):1063–70. doi:10.1037/0022-3514.54.6.1063.

Pejtersen JH, Burr H, Hannerz H, Fishta A, Eller NH. Update on work-related psychosocial factors and the development of ischemic heart disease. Cardiol Rev. 2015; 23(2):94–98. doi:10.1097/crd.0000000000000033.

Theorell T, Hammarström A, Aronsson G, Bendz LT, Grape T, Hogstedt C, Marteinsdottir I, Skoog I, Hall C. A systematic review including meta-analysis of work environment and depressive symptoms. BMC Public Health. 2015; 15(738). http://dx.doi.org/10.1186/s12889-015-1954-4.

Burr H, Formazin M, Pohrt A. Methodological and conceptual issues regarding occupational psychosocial coronary heart disease epidemiology. Scand J Work Environ Health. 2016; 42(3):251–55. doi:10.5271/sjweh.3557.

PubMed   Google Scholar  

Alberdi A, Aztiria A, Basarab A. Towards an automatic early stress recognition system for office environments based on multimodal measurements. J Biomed Inform. 2016; 59:49–75. doi:10.1016/j.jbi.2015.11.007.

Baethge A, Vahle-Hinz T, Rigotti T. Coworker support and its relationship to allostasis during a workday: A diary study on trajectories of heart rate variability during work. J Appl Psychol. 2020; 105(5):506–26. doi:10.1037/apl0000445.

Almeida DM, Davis KD, Lee S, Lawson KM, Walter KN, Moen P. Supervisor support buffers daily psychological and physiological reactivity to work-to-family conflict. J Marriage Fam. 2016; 78(1):165–79. doi:10.1111/jomf.12252.

Martikainen P, Bartley M, Lahelma E. Psychosocial determinants of health in social epidemiology. Int J Epidemiol. 2002; 31(6):1091–93. doi:10.1093/ije/31.6.1091.

Rugulies R. What is a psychosocial work environment?. Scand J Work Environ Health. 2019; 45(1):1–6. doi:10.5271/sjweh.3792.

Kim TJ, von dem Knesebeck O. Is an insecure job better for health than having no job at all? a systematic review of studies investigating the health-related risks of both job insecurity and unemployment. BMC Public Health. 2015; 15(1). http://dx.doi.org/10.1186/s12889-015-2313-1.

Steptoe A, Kivimäki M. Stress and cardiovascular disease. Nat Rev Cardiol. 2012; 9(6):360–70. doi:10.1038/nrcardio.2012.45.

Steptoe A, Kivimäki M. Stress and cardiovascular disease: An update on current knowledge. Annu Rev Public Health. 2013; 34(1):337–54. doi:10.1146/annurev-publhealth-031912-114452.

Belkic K, Landsbergis PA, Schnall PL, Baker D. Is job strain a major source of cardiovascular disease risk?Scand J Work Environ Health. 2004; 30(2):85–128. doi:10.5271/sjweh.769.

Kivimäki M, Steptoe A. Effects of stress on the development and progression of cardiovascular disease. Nat Rev Cardiol. 2018; 15(4):215–29. doi:10.1038/nrcardio.2017.189.

Kivimäki M, Virtanen M, Elovainio M, Kouvonen A, Väänänen A, Vahtera J. Work stress in the etiology of coronary heart disease—a meta-analysis. Scand J Work Environ Health. 2006; 32(6):431–42. doi:10.5271/sjweh.1049.

Siegrist J, Li J. Associations of extrinsic and intrinsic components of work stress with health: A systematic review of evidence on the effort-reward imbalance model. Int J Environ Res Public Health. 2016; 13(4):432. doi:10.3390/ijerph13040432.

Wasserstein RL, Lazar NA. The ASA statement on p-values: Context, process, and purpose. Amer Statist. 2016; 70(2):129–33. doi:10.1080/00031305.2016.1154108.

Louch G, O’Hara J, Gardner P, O’Connor DB. A daily diary approach to the examination of chronic stress, daily hassles and safety perceptions in hospital nursing. Int J Behav Med. 2017; 24(6):946–56. doi:10.1007/s12529-017-9655-2.

Beckers DGJ, van Hooff MLM, van der Linden D, Kompier MAJ, Taris TW, Geurts SAE. A diary study to open up the black box of overtime work among university faculty members. Scand J Work Environ Health. 2008; 34(3):213–23. doi:10.5271/sjweh.1226.

Rutledge T, Stucky E, Dollarhide A, Shively M, Jain S, Wolfson T, Weinger MB, Dresselhaus T. A real-time assessment of work stress in physicians and nurses. Health Psychol. 2009; 28(2):194–200. doi:10.1037/a0013145.

Ilies R, Johnson MD, Judge TA, Keeney J. A within-individual study of interpersonal conflict as a work stressor: Dispositional and situational moderators. J Organ Behav. 2011; 32(1):44–64. doi:10.1002/job.677. First published: 22 December 2010.

Yeh Y-JY, Ma T-N, Pan S-Y, Chuang P-J, Jhuang Y-H. Assessing potential effects of daily cross-domain usage of information and communication technologies. J Soc Psychol. 2019; 160(4):465–78. doi:10.1080/00224545.2019.1680943.

Zhou L, Wang M, Chang C-H, Liu S, Zhan Y, Shi J. Commuting stress process and self-regulation at work: Moderating roles of daily task significance, family interference with work, and commuting means efficacy. Pers Psychol. 2017; 70(4):891–922. doi:10.1111/peps.12219.

Buunk BP, Verhoeven K. Companionship and support at work: A microanalysis of the stress-reducing features of social interaction. Basic Appl Soc Psych. 1991; 12(3):243–58. http://dx.doi.org/10.1207/s15324834basp1203_1.

Dudenhöffer S, Dormann C. Customer-related social stressors and service providers’ affective reactions. J Organ Behav. 2012; 34(4):520–39. doi:10.1002/job.1826.

Rodrigues S, Kaiseler M, Queirós C, Basto-Pereira M. Daily stress and coping among emergency response officers: a case study. Int J Emerg Serv. 2017; 6(2):122–33. doi:10.1108/ijes-10-2016-0019.

Beattie L, Griffin B. Day-level fluctuations in stress and engagement in response to workplace incivility: A diary study. Work Stress. 2014:1–19. doi:10.1080/02678373.2014.898712.

Diebig M, Bormann KC, Rowold J. Day-level transformational leadership and followers’ daily level of stress: A moderated mediation model of team cooperation, role conflict, and type of communication. Eur J Work Organ Psy. 2017; 26(2):234–49. doi:10.1080/1359432x.2016.1250741.

Webster JR, Adams GA, Maranto CL, Beehr TA. “dirty” workplace politics and well-being. Psychol Women Quart. 2018; 42(3):361–77. doi:10.1177/0361684318769909.

Kamarck TW, Shiffman SM, Smithline L, Goodie JL, Paty JA, Gnys M, Jong JY-K. Effects of task strain, social conflict, and emotional activation on ambulatory cardiovascular activity: Daily life consequences of recurring stress in a multiethnic adult sample. Health Psychol. 1998; 17(1):17–29. doi:10.1037/0278-6133.17.1.17.

Albrecht SL, Anglim J. Employee engagement and emotional exhaustion of fly-in-fly-out workers: A diary study. Aust J Psychol. 2018; 70(1):66–75. doi:10.1111/ajpy.12155.

Breevaart K, Bakker AB, Derks D, van Vuuren TCV. Engagement during demanding workdays: A diary study on energy gained from off-job activities. Int J Stress Manage. 2020; 27(1):45–52. doi:10.1037/str0000127.

Karlsson K, Niemelä P, Jonsson A. Heart rate as a marker of stress in ambulance personnel: A pilot study of the body’s response to the ambulance alarm. Prehospital Disaster Med. 2011; 26(1):21–26. doi:10.1017/s1049023x10000129.

Fernández-Castro J, Martínez-Zaragoza F, Rovira T, Edo S, Solanes-Puchol Á, Martín-del-Río B, García-Sierra R, Benavides-Gil G, Doval E. How does emotional exhaustion influence work stress? relationships between stressor appraisals, hedonic tone, and fatigue in nurses’ daily tasks: A longitudinal cohort study. Int J Nurs Stud. 2017; 75:43–50. doi:10.1016/j.ijnurstu.2017.07.002.

Klumb PL, Voelkle MC, Siegler S. How negative social interactions at work seep into the home: A prosocial and an antisocial pathway. J Organ Behav. 2017; 38(5):629–49. doi:10.1002/job.2154.

Pereira D, Semmer NK, Elfering A. Illegitimate tasks and sleep quality: An ambulatory study. Stress Health. 2014; 30(3):209–21. doi:10.1002/smi.2599.

Stucky ER, Dresselhaus TR, Dollarhide A, Shively M, Maynard G, Jain S, Wolfson T, Weinger MB, Rutledge T. Intern to attending: Assessing stress among physicians. Acad Med. 2009; 84(2):251–57. doi:10.1097/acm.0b013e3181938aad.

Baethge A, Rigotti T. Interruptions to workflow: Their relationship with irritation and satisfaction with performance, and the mediating roles of time pressure and mental demands. Work Stress. 2013; 27(1):43–63. doi:10.1080/02678373.2013.761783.

Ferdous R, Osmani V, Marquez JB, Mayora O. Investigating correlation between verbal interactions and perceived stress. In: 2015 37th Annual International Conference of the IEEE Engineering in Medicine and Biology Society (EMBC). IEEE: 2015. p. 1612–15. doi:10.1109/embc.2015.7318683.

Matta FK, Scott BA, Colquitt JA, Koopman J, Passantino LG. Is consistently unfair better than sporadically fair? an investigation of justice variability and stress. Acad Manage J. 2017; 60(2):743–70. doi:10.5465/amj.2014.0455.

Reicherts M, Pihet S. Job newcomers coping with stressful situations: A micro-analysis of adequate coping and well-being. Swiss J Psychol. 2000; 59(4):303–16. doi:10.1024//1421-0185.59.4.303.

Gervais RL. Menstruation as a work stressor: Evidence and interventions In: Gervais R. L., Millear P. M., editors. Exploring Resources, Life-Balance and Well-Being of Women Who Work in a Global Context. Springer: 2016. p. 201–18. Chap. 12. http://dx.doi.org/10.1007/978-3-319-31736-6_1 .

Boudreaux E, Jones GN, Mandry C, Brantley PJ. Patient care and daily stress among emergency medical technicians. Prehospital Disaster Med. 1996; 11(3):188–93. doi:10.1017/s1049023x0004293x.

Shively M, Rutledge T, Rose BA, Graham P, Long R, Stucky E, Weinger MB, Dresselhaus T. Real-time assessment of nurse work environment and stress. J Healthc Qual. 2011; 33(1):39–48. doi:10.1111/j.1945-1474.2010.00093.x.

Baethge A, Deci N, Dettmers J, Rigotti T. “some days won’t end ever”: Working faster and longer as a boundary condition for challenge versus hindrance effects of time pressure. J Occup Health Psychol. 2019; 24(3):322–32. doi:10.1037/ocp0000121.

Buunk BP, Peeters MCW. Stress at work, social support and companionship: Towards an event-contingent recording approach. Work Stress. 1994; 8(2):177–90. doi:10.1080/02678379408259988.

Weenk M, Alken APB, Engelen LJLPG, Bredie SJH, van de Belt TH, van Goor H. Stress measurement in surgeons and residents using a smart patch. Am J Surg. 2018; 216(2):361–68. doi:10.1016/j.amjsurg.2017.05.015.

Steptoe A. Stress, social support and cardiovascular activity over the working day. Int J Psychophysiol. 2000; 37(3):299–308. doi:10.1016/s0167-8760(00)00109-4.

Johnston D, Bell C, Jones M, Farquharson B, Allan J, Schofield P, Ricketts I, Johnston M. Stressors, appraisal of stressors, experienced stress and cardiac response: A real-time, real-life investigation of work stress in nurses. Ann Behav Med. 2016; 50(2):187–97. doi:10.1007/s12160-015-9746-8.

van Hooff MLM. The daily commute from work to home: Examining employees’ experiences in relation to their recovery status. Stress Health. 2015; 31(2):124–37. doi:10.1002/smi.2534.

Diebig M, Bormann KC. The dynamic relationship between laissez-faire leadership and day-level stress: A role theory perspective. Ger J Hum Resour Manag. 2020; 34(3):324–44. doi:10.1177/2397002219900177.

Wood S, Michaelides G, Totterdell P. The impact of fluctuating workloads on well-being and the mediating role of work-nonwork interference in this relationship. J Occup Health Psychol. 2013; 18(1):106–19. doi:10.1037/a0031067.

Levin S, France DJ, Hemphill R, Jones I, Chen KY, Rickard D, Makowski R, Aronsky D. Tracking workload in the emergency department. Hum Factors. 2006; 48(3):526–39. doi:10.1518/001872006778606903.

Tadić M, Bakker AB, Oerlemans WGM. Work happiness among teachers: A day reconstruction study on the role of self-concordance. J School Psychol. 2013; 51(6):735–50. doi:10.1016/j.jsp.2013.07.002.

Elfering A, Semmer NK, Grebner S. Work stress and patient safety: Observer-rated work stressors as predictors of characteristics of safety-related events reported by young nurses. Ergonomics. 2006; 49(5-6):457–69. doi:10.1080/00140130600568451.

Almeida DM, Davis KD. Workplace flexibility and daily stress processes in hotel employees and their children. Ann Am Acad Politi Soc Sci. 2011; 638(1):123–40. doi:10.1177/0002716211415608.

Download references

Acknowledgments

Not applicable.

This work was supported by the Research Foundation – Flanders, Belgium (FWO) under Grant (project no. G.0318.18N); and the Slovenian Research Agency (ARRS) under Grant (project ref. N2-0081).

Author information

Junoš Lukan and Larissa Bolliger contributed equally to this work.

Authors and Affiliations

Department of Intelligent Systems, Jožef Stefan Institute, Jamova cesta 39, Ljubljana, 1000, Slovenia

Junoš Lukan & Mitja Luštrek

Jožef Stefan Postgraduate School, Jamova cesta 39, Ljubljana, 1000, Slovenia

Department of Public Health and Primary Care, Ghent University, C. Heymanslaan 10, Ghent, 9000, Belgium

Larissa Bolliger, Dirk De Bacquer & Els Clays

Knowledge Centre for Health Ghent, Ghent University, C. Heymanslaan 10, Ghent, 9000, Belgium

Nele S. Pauwels

Ghent University Hospital, C. Heymanslaan 10, Ghent, 9000, Belgium

You can also search for this author in PubMed   Google Scholar

Contributions

The following authors all contributed to this work: Junoš Lukan (JL), Larissa Bolliger (LB), Nele S. Pauwels (NP), Mitja Luštrek (ML), Dirk De Bacquer (DB), Els Clays (EC). The roles they shared according to the CRediT designations are as follows. Conceptualization: JL, LB, ML, EC Data curation: JL, LB Formal analysis: JL, LB Funding acquisition: ML, DB, EC Investigation: JL, LB Methodology: Not applicable Project administration: JL, LB, ML, EC Resources: NP Software: Not applicable Supervision: NP, ML, DB, EC Validation: Not applicable Visualization: JL Writing - original draft: JL, LB Writing - review & editing: JL, LB, NP, ML, EC. The authors read and approved the final manuscript.

Corresponding author

Correspondence to Larissa Bolliger .

Ethics declarations

Ethics approval and consent to participate, consent for publication, competing interests.

The authors declare that they have 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

Additional file 1.

Tables 1 and 2 show work environment risk factors and stress outcomes, respectively, together with the tools used to measure them in original studies. They are ordered in broader categories according to the 6th European Working Conditions Survey [ 29 ] and with respect to the stress model of Ice and James [ 30 ].

Additional file 2

Correlations were extracted from studies that reported them. They are listed in a spreadsheet to enable replication of analyses. These correlations were counted to produce Fig.  2 in the main body of text.

Additional file 3

The full search strategies with indexing terms and free text words for all the databases searched: PubMed, Embase, Web of Science, Scopus, CINAHL, ERIC, and PsycArticles.

Additional file 4

The PRISMA Checklist noting the page numbers of all necessary review items.

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.

Lukan, J., Bolliger, L., Pauwels, N.S. et al. Work environment risk factors causing day-to-day stress in occupational settings: a systematic review. BMC Public Health 22 , 240 (2022). https://doi.org/10.1186/s12889-021-12354-8

Download citation

Received : 09 April 2021

Accepted : 29 November 2021

Published : 05 February 2022

DOI : https://doi.org/10.1186/s12889-021-12354-8

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

  • Day-to-day stress
  • Ecological Momentary Assessment (EMA)
  • Work environment risk factors
  • Stress outcomes
  • Systematic literature review

BMC Public Health

ISSN: 1471-2458

research paper on stress in the workplace

  • - Google Chrome

Intended for healthcare professionals

  • Access provided by Google Indexer
  • My email alerts
  • BMA member login
  • Username * Password * Forgot your log in details? Need to activate BMA Member Log In Log in via OpenAthens Log in via your institution

Home

Search form

  • Advanced search
  • Search responses
  • Search blogs

Stress at work

  • Related content
  • Peer review

This article has a correction. Please see:

  • Errata - June 29, 2017
  • Thomas Despréaux , chief resident 1 2 3 ,
  • Olivier Saint-Lary , general practitioner , senior lecturer 4 5 ,
  • Florence Danzin , psychiatrist 1 6 ,
  • Alexis Descatha , occupational/emergency practitioner , professor 1 2 3
  • 1 Occupational health unit, University hospital of Poincaré site, Garches, France
  • 2 Versailles St-Quentin University, Versailles, France
  • 3 CESP, U 1018 Inserm, Villejuif, France
  • 4 Versailles Saint-Quentin en Yvelines, Faculty of Health sciences Simone Veil, Department of Family Medicine, Montigny le Bretonneux, France
  • 5 Université Paris-Saclay, University Paris-Sud, Villejuif, France
  • 6 Charcot Psychiatric Hospital, France
  • Correspondence to O Saint-Lary olivier.saint-lary{at}uvsq.fr

What you need to know

Long working hours and strain at work contribute to stress, ill health, and increased risk of cardiovascular diseases, diabetes, and mental illnesses

Explore occupational factors such as an imbalance between effort and reward, work overload, bullying, and job insecurity

Workplace interventions, a short period of leave from work, and psychological treatment can be considered, alongside regular follow-up to assess how the patient is coping

A 55 year old senior executive presents with low back pain. He appears anxious. A reorganisation within his company has increased his workload and he has been working more hours but receiving no recognition from management. Last week he felt humiliated by a colleague. Since then he has not been able to sleep for more than a couple of hours each day.

Stress accounts for more than a third of all cases of work related ill health and almost half of all working days lost due to illness. 1 Internationally, systematic reviews and meta-analysis of observational data suggest that job strain and poorly functioning work environments are associated with the development of depressive symptoms. 2 3 4 A longitudinal cohort study from Norway found workplace bullying to be associated with subsequent suicidal ideation. 5 Long working hours are also associated with increased risk of stroke, heart disease, 6 and diabetes, 7 and poor lifestyle including inactivity, 7 smoking, 7 and risky alcohol use. 8

Patients might present with unexplained somatic symptoms, such as odd aches and pains, palpitations, loss of appetite, and loss of sleep. 9 10 Explore their symptoms and discuss any contributing factors in their work and personal life. The consultation can be long and difficult, as the patient might not volunteer all the information or draw the association with work stress. The objective of this first consultation is to perform a quick risk assessment and explore factors in the patient’s job that are contributing to stress.

What you should cover

The following questions are based on systematic reviews, and the experiences of clinicians and patients.

• the nature and duration of the patient’s presenting symptoms

• associated depressive symptoms, such as

o feeling down, low, or sad

o loss of interest in activities

o tiring easily

o lack of concentration

o changes in sleep and appetite

• feelings of hopelessness, (eg, a belief that the situation cannot improve) 11

• occupation, working environment, and stressors at work (box 1)

• the chronology of events, how the patient has coped so far, and if things have changed recently in their workplace. Typically, three phases are described 13 :

an initial (“serene”) phase, where the patient reports no particular difficulty

a “problem” phase, when obstacles and conflicts gradually appear and the patient tries to deal with the situation

a “crisis” phase, where s/he comes to see you

• protective factors for severity of outcome include a supportive family environment and financial wellbeing. Aggravating factors are familial isolation, being a single parent with young children, having financial difficulties, or being bound by a particular type of employment contract that forces the patient to stay in the same job. The latter can delay diagnosis, and limit the range of remedial options available.

• thoughts of ending their life or causing harm to themselves or others

• other medical illnesses, including diabetes, hypertension, cardiovascular events, or psychiatric disorders

• smoking, alcohol, and drug abuse

• family history of depression or mental disorders, which could increase the risk of depression and suicide

Box 1: Occupational factors for stress 2 12 13

Conflict of values (being asked to do a poor quality job or cut costs for a person who likes to keep high standards in their work)

Feeling insufficiently rewarded compared with the person’s assessment of their efforts (“effort-reward imbalance”)

Inability to make decisions about when or how to stop work

Lack of support from colleagues and management

Isolation at work (no cooperation between teams)

Work overload (working after hours) or insufficient workload (nothing to do)

Discrimination, humiliation, violence, bullying, and harassment at work

Cases of work related stress in the same company

Company situation in terms of finances, organisational changes, and employee turnover

Job insecurity, temporary employment status

Patients come to their doctor primarily to address their symptoms, but some will also want assistance and advice on how to cope with the situation at work.

Examination

Assess general appearance and look for signs of psychomotor agitation such as restlessness, rapid talking, and racing thoughts, or of psychomotor retardation such as apparent exhaustion and visible slowing of physical activity. These might indicate a mental illness or organic cause, such as a thyroid disorder.

Perform a quick general examination to look for fever, tachycardia, hypertension, and signs of thyroid disorder (which can be a differential diagnosis). Examine thoroughly for reported pain, though somatisation is likely.

What you should do

Investigation and management of physical and mental health diagnoses —Offer usual management of conditions such as depression. Consider immediate referral to psychiatry if the patient describes suicidal or aggressive thoughts or intentions.

Make the connection between the patient’s experience and work stress —For patients with work related stress and a variety of symptoms, acknowledge their situation and validate their feelings with a phrase such as, “I understand that you are suffering and that this feeling is arising from a stressful work environment.”

Offer a supportive setting to discuss and make progress in dealing with work stress —High quality evidence and guidelines for interventions to manage work related adjustment issues and stress are lacking. 14 Cognitive therapy, stepwise reintegration planning, and relaxation training can all be considered. 15 16 Therapy needs to be supportive, active, flexible, goal directed, and time bound. 10 12 14

Consider offering a second appointment—for example, if there is too much to cover. You might suggest that the patient brings a family member to the next appointment for support.

In the interim, you might ask the patient to reflect on their job and personal situation, and possibly to write a short description of their problems at work, the chronology of these problems, and their relationship to the patient’s symptoms. In our experience, some patients find this helps them reflect on the events, and it can help you understand their situation better. This will help to initiate discussion on strategies that the patient might employ to navigate their workspace going forward. Making contact with the workplace to modify work or reduce workload in collaboration with the employer can be helpful. Discuss whether the occupational health services or human resources division at the patient’s company could be involved. In some circumstances, patients might wish to seek compensation or take legal action. Explore if these are important for your patient and direct them to appropriate agencies or lawyers who can help with these matters.

Consider whether the patient wants or might benefit from time away from work including a “sick note.”

Schedule a follow-up visit to assess how the patient is coping with symptoms and workplace issues, and modify the approach accordingly.

Education into practice

What factors would you typically explore in the patient’s history to understand their working environment and stress? Does this article offer you ideas on how to do so differently?

Sometimes, asking the patient to write down their problems at work, the times at which the problems occurred, and the patient’s symptoms, is helpful. Are there ways in which you might consider using this or other techniques to help patients better organise their thoughts or understand them yourself?

Do you offer a second appointment, if there is too much to cover, or if the patient wishes to include a friend or family member?

In difficult cases, do you work in collaboration with mental health professionals as well as occupational health professionals?

How patients were involved in the creation of this article

Patients in our practice reported a need to rethink what they had experienced at work and to share this in writing. This helped them identify and clearly communicate the chronology of events. Based on their feedback we recommend encouraging patients to write about their work environment and factors contributing to stress, though this need not be mandatory.

A patient reviewed this article and attested that writing a two page memorandum would have been enormously helpful to identify problems at work and how they had escalated over time, and to come to terms with the situation.

We would like to thank Richard Carter for helping us to improve the language of this document.

We have read and understood BMJ policy on declaration of interests and declare that we have no competing interests.

Patient consent obtained.

This is an Open Access article distributed in accordance with the Creative Commons Attribution Non Commercial (CC BY-NC 4.0) license, which permits others to distribute, remix, adapt, build upon this work non-commercially, and license their derivative works on different terms, provided the original work is properly cited and the use is non-commercial. See: http://creativecommons.org/licenses/by-nc/4.0/ .

  • ↵ Statistics—Work related stress, anxiety and depression statistics in Great Britain (GB). http://www.hse.gov.uk/statistics/causdis/stress
  • ↵ Theorell T, Hammarström A, Aronsson G, et al. A systematic review including meta-analysis of work environment and depressive symptoms. BMC Public Health 2015 ; 15 : 738 . doi:10.1186/s12889-015-1954-4 . pmid:26232123 . OpenUrl
  • ↵ Rugulies R, Aust B, Madsen IE. Effort-reward imbalance at work and risk of depressive disorders. A systematic review and meta-analysis of prospective cohort studies. Scand J Work Environ Health 2017 ; 17 : 3632 . doi:10.5271/sjweh.3632 . pmid:28306759 . OpenUrl
  • ↵ Harvey SB, Modini M, Joyce S, et al. Can work make you mentally ill? A systematic meta-review of work-related risk factors for common mental health problems. Occup Environ Med 2017 ; 74 : 301 - 10 . doi:10.1136/oemed-2016-104015 . pmid:28108676 . OpenUrl
  • ↵ Nielsen MB, Einarsen S, Notelaers G, Nielsen GH. Does exposure to bullying behaviors at the workplace contribute to later suicidal ideation? A three-wave longitudinal study. Scand J Work Environ Health 2016 ; 42 : 246 - 50 . doi:10.5271/sjweh.3554 . pmid:27135593 . OpenUrl
  • ↵ Kivimäki M, Jokela M, Nyberg ST, et al. Long working hours and risk of coronary heart disease and stroke: a systematic review and meta-analysis of published and unpublished data for 603 838 individuals. Lancet 2015;386:1739-46. doi:10.1016/S0140-6736(15)60295-1
  • ↵ Nyberg ST, Fransson EI, Heikkilä K, et al. Job Strain and Cardiovascular Disease Risk Factors: Meta-Analysis of Individual-Participant Data from 47 000 Men and Women. Testa L, ed. PLoS ONE 2013;8:e67323. doi:10.1371/journal.pone.0067323
  • ↵ Virtanen M, Jokela M, Nyberg ST, et al. Long working hours and alcohol use: systematic review and meta-analysis of published studies and unpublished individual participant data. BMJ 2015;350:g7772. doi:10.1136/bmj.g7772
  • ↵ American Psychiatric Association. Diagnostic and Statistical Manual of Mental Disorders. 5th ed (DSM-5). American Psychiatric Publishing 2013.
  • ↵ van der Klink JJL, van Dijk FJH. Dutch practice guidelines for managing adjustment disorders in occupational and primary health care. Scand J Work Environ Health 2003 ; 29 : 478 - 87 . doi:10.5271/sjweh.756 pmid:14712856 . OpenUrl
  • ↵ Fraser L, Burnell M, Salter LC, et al. Identifying hopelessness in population research: a validation study of two brief measures of hopelessness. BMJ Open 2014 ; 4 : e005093 . doi:10.1136/bmjopen-2014-005093 . pmid:24879829 . OpenUrl
  • ↵ Nieuwenhuijsen K, Bruinvels D, Frings-Dresen M. Psychosocial work environment and stress-related disorders, a systematic review. Occup Med (Lond) 2010 ; 60 : 277 - 86 . doi:10.1093/occmed/kqq081 . pmid:20511268 . OpenUrl
  • ↵ Mediouni Z, Garrabé H, Jaworski F, et al. Initial evaluation of patients reporting a work-related stress or bullying. J Occup Environ Med 2012 ; 54 : 1439 - 40 . doi:10.1097/JOM.0b013e31827942e0 . pmid:23222476 . OpenUrl
  • ↵ Joosen MCW, Brouwers EPM, van Beurden KM, et al. An international comparison of occupational health guidelines for the management of mental disorders and stress-related psychological symptoms. Occup Environ Med 2015 ; 72 : 313 - 22 . doi:10.1136/oemed-2013-101626 . pmid:25406476 . OpenUrl
  • ↵ West CP, Dyrbye LN, Erwin PJ, Shanafelt TD. Interventions to prevent and reduce physician burnout: a systematic review and meta-analysis. Lancet 2016 ; 388 : 2272 - 81 . doi:10.1016/S0140-6736(16)31279-X . pmid:27692469 . OpenUrl
  • ↵ Van den Broeck K, Remmen R, Vanmeerbeek M, Destoop M, Dom G. Collaborative care regarding major depressed patients: A review of guidelines and current practices. J Affect Disord 2016 ; 200 : 189 - 203 . doi:10.1016/j.jad.2016.04.044 . pmid:27136418 . OpenUrl

research paper on stress in the workplace

Academia.edu no longer supports Internet Explorer.

To browse Academia.edu and the wider internet faster and more securely, please take a few seconds to  upgrade your browser .

Enter the email address you signed up with and we'll email you a reset link.

  • We're Hiring!
  • Help Center

paper cover thumbnail

Stress at the Workplace and Its Impacts on Productivity: A Systematic Review from Industrial Engineering, Management, and Medical Perspective

Profile image of Elkana Timotius

2022, Industrial Engineering & Management Systems

In every fast-paced surrounding, stress is present in every life aspect, including at the workplace. It is a deeply personal experience, with various stressors affecting every individual differently. This study assessed the past and present workplace stress-related information and analyzed its impact on productivity. It primarily concentrates on the field's philosophical principles, while providing a collection of directions for future study as well. This study was formed in the statement of PRISMA (Preferred Reporting Items for Systematic Review and Meta-Analysis). The impact of stress at the workplace on the employee's productivity was observed in the cohort and cross-sectional studies from the perspective of industrial engineering, management, and medicine. Four eligible studies were qualitatively assessed from 2,642 identified literature through four databases (Cochrane, Science Direct, Scielo, and PubMed) using keywords stress, impact, productivity, industrial engineering, management, and medicine. The study was convinced that stress at the workplace contributes to worsening relationships at home, worsening relationships between superiors and subordinates as well as contracting diseases. It has a potential negative impact on productivity. Furthermore, the work environment plays a significant contribution in inducing workplace stress because of human physiologic response. Noxious stress is detrimental to the human body, especially if maintained in the long run. Therefore, stress management is imperative before it is too late.

Related Papers

Ebenezer Ofosuhene

This review of the literature gives information about work stress, factors in the working environment that cause stressful situations and negative health consequences of the workplace stress. Stressors are pointed out in details that lead to stress at the workplace. Approaches to the stress are explained and most famous models of the stress are assessed critically in this review. This article highlights the work stress and its adverse effects on the physical and mental health of an employee. Finally, recommendations for future research are given and areas are highlighted where there is need of more empirical research.

research paper on stress in the workplace

Scandinavian Journal of Work, Environment &amp; Health

Umesh Maiya

Stress is much in the news at present but it isn&#39;t a new problem. Pressure is part and parcel of all work and helps to keep us motivated. But excessive pressure can lead to stress which undermines performance, is costly to employers and can make people ill. Research reveals that many working days are lost to stress, depression and anxiety. Work-related stress costs a huge burden to the society. Stress takes many forms as well as leading to anxiety and depression it can have a significant impact on an employee&#39;s physical health. Research links stress to heart disease, back pain, headaches, gastrointestinal disturbances and alcohol and drug dependency. Individuals are more willing to admit that they are suffering from stress if they can expect to be dealt with sympathetically. In some cases good counseling may be all that is needed. This paper aims at studying the stressors that affects an individual at work, to examine the effects of stress and suitable measures which employe...

AAOHN Journal

Bonita Long

Michael Murray

Rex Journal

In today's age of automation, advanced technology & high competition, man has great dreams of a luxurious living & enjoys at the thought of experiencing it. It is a well-accepted fact that every human being is an individual with his own unique characteristics & ways of responding & behaving. These various ways of responding & behaving can be either positive or negative & these can make one's life a happy or a miserable one. These facts are true for every individual in every sphere of life. In today's fast moving world every individual strives hard to achieve their dreams & the best of luxurious living but faces stress in the process of doing so. The present study will bring to light the stress level, sources of stress & stress management strategies.

Littera Scripta

Andrea Bencsik

Erick Onsongo

Over the past few decades the stress had become a growing dilemma in organization and cause unfavorable effects on job performance. Stress is a universal element that affects employees worldwide. There are many barriers that affecting the employees in the workplace. Work stress often affects the employees in the workplace, where each employee will feel it at least once in their workplace. Work stress is a real life problem that not only affects the organization, but the employees mainly become victims of stress. stress become a familiar element in organization and nowadays the workplace become more complicated, which bring more negative impacts to the employees compared to positive impacts. Stress among workers is greater than before which also effect on the whole performance of the employees. Stress which occurred in workplace declared as harmful to physical and emotional responses that happen within a human being when the requirement of the job do not match the employees‟ capacity...

Norliyana Zakaria

The systematic review in this study is looking at the causes of stress in the workplace. The authors will identify the most leading causes of work-related stress, as well as their impacts on employees’ physical well-being, mental health and work performance. In addition, the stress factors arising from work environment or atmosphere, workplace relationships, organizational culture, career growth, role conflict, and work-life balance are highlighted. Finally, the authors discuss on several stress coping approaches, namely getting social support, attending stress coping or stress management programmes, upgrading work environment, and reforming organizational culture.

RELATED PAPERS

2020 IEEE PES Innovative Smart Grid Technologies Europe (ISGT-Europe)

Mart Van Der Meijden

Ratih Fitrianti

EJVES Extra

Charles McCollum

Revista Chilena De Derecho

Hernan Corral Talciani

Osvaldo Gomes

arXiv: Methodology

Frederic Bertrand

Jurnal Ilmiah Betrik

khandava mulyadien

Proceedings of the International Symposium “Mesoscale and Submesoscale Processes in the Hydrosphere and Atmosphere” (MSP-2018), 30 October – 02 November 2018, Moscow / [Ed. by: Zatsepin A.G., Ginzburg A.I., Kostianoy A.G., Sviridov S.A.]

Vasiliy Melnikov

RYSZARD MIKLASZEWSKI

Rini Punathil

Transgenic Research

Venkata Tavva

Remote Sensing

Kamila Pawłuszek

European Journal of Clinical Nutrition

Maria ortega

Southeast European Review of Business and Economics

Ivana Panova

Scientific reports

Pradip Dutta

Heather Hoffman

Nanotechnology

sandro santucci

Dominic Furniss

Afrah Chowdhury

JBRA assisted reproduction

Tracy Irani

Gerhard Tauberschmidt

Journal of physics

George Nikiforidis

Natural Hazards and Earth System Sciences Discussions

RePEc: Research Papers in Economics

Paolo Sestito

RELATED TOPICS

  •   We're Hiring!
  •   Help Center
  • Find new research papers in:
  • Health Sciences
  • Earth Sciences
  • Cognitive Science
  • Mathematics
  • Computer Science
  • Academia ©2024

Stress Factors in the Workplace Research Paper

Introduction.

Stress refers to how the body responds to any change. A stress factor or stressor is a situation or pressure that results in that experience. Normally, people think of stress factors as negative such as exhausting work schedules, difficult clients to work with, other challenges, and family relations. Contrary to this notion, anything that places high demands on a person can be stressful. How individuals handle such situations determines their impact on their lives. Long-term exposure to stress factors can be a great concern. The body starts to react negatively toward the stress, which can result in poor health with time. In work settings, this can influence the productivity of an employee. Currently, in the United States, there is a shortage of nurses in the healthcare sector. The morale of healthcare professionals has been reduced, and industry leaders must find ways to help the workers. This paper aims to address the stress factors in the workplace and programs organizations could implement to help employee stress.

Stress Factors

Regardless of the sector, co-workers can create a toxic culture or environment within the work setting. Most of the time, such surroundings are characterized by a lack of collaboration and cooperation among employees. The choice to work from home during the pandemic may have felt great for those in organizations with this type of culture (Chachula & Ahmad, 2022). However, there are still such occurrences, even for individuals working remotely. Toxic traits by colleagues do not simply stop since they are not in person. They can even follow someone or observe through virtual communication, including zoom.

A toxic workplace consists of an environment where dysfunction is found. Such surroundings impact employees by making them feel disengaged and demoralized. It can result in high stress levels, depression, fatigue, anxiety, and burnout. The report shows that in healthcare facilities or organizations, nurses become unmotivated to offer proper care services to the patients, and thus, their productivity lowers (Chachula & Ahmad, 2022). Cases of negligence are twenty percent higher in institutions with toxic cultures (Chachula & Ahmad, 2022). In a sector that directly deals with people’s lives, finding a solution to the problem is important.

Bullying and harassment

In extreme instances, some colleagues can be demanding and manipulative and controlling, which is equivalent to bullying a person or harassing them. Study shows that 29% of individuals have been bullied in work settings, and 72% reported that the cases were performed by older employees of particular organizations (Chachula & Ahmad, 2022). Stress and poor health can become part of the everyday life of individuals being harassed. Another research claims that workers believe that many organizational leaders overlook reports of bullying and harassment (Díaz McConnell et al., 2022). By doing this, the managers and bosses are enabling bad behavior that is ruining the wellbeing of their younger employees.

Poor Communication

Ambiguousness and inadequate or misleading communication at workplaces can result in unidentified issues. This is usually followed by the inability of a fellow employee to listen. A study revealed that this is the most common cause of stress at work (Díaz McConnell et al., 2022). Poor communication can be in several forms, but the results or outcomes are the same. This mostly happens to new employees at an organization (Díaz McConnell et al., 2022). For instance, when nurses are first employed and given specific roles to perform at a healthcare facility, it is the role of the more experienced workers to help by giving them directions and needed information. However, in some instances, the opposite happens and leads to anxiety as the employees feel as if they will be forced to leave work due to failure to perform their tasks. As this continues over time, it causes them to develop stress.

Relationships at Work

Employees spend more time with each other than with friends or family, and the choice of who their colleagues are is determined by the leaders of an organization. Individuals are hired for their capabilities and not their matching or compatibility with others, which might lead to friction between peers. Even though most workplaces can be supportive, it is not always the case. People can be susceptible to unhealthy competition or a trait such as jealousy (Díaz McConnell et al., 2022). If this is unchecked, it can grow into intimidation, bullying, or gossiping. Going to a work environment where colleagues’ behavior undermines a person is highly stressful. It is not always easy to identify, but a great leader should be able to notice the tension between members of the staff.

Stress with Patients

A nurse is highly likely to develop stress after the death of a patient they connected with through offering them care. Nurses have more contact with their clients than other healthcare professionals. In some aspects of hospital care, an individual needs hospitalization multiple times, or it is prolonged due to the condition. This enables the nurses to know the patients and families well. The more often the sides meet, the greater the bond established between them. Dying is an inseparable as well as integral phase of life. The phenomenon has always followed human existence (Jordan et al., 2022). The atmosphere of death can be associated with personal fears and feelings, which make the experience depressing and distressing.

Dying is a challenging time in a sick individual’s life as it is filled with anxiety for the person themselves and those around them. In the work settings of a hospital or healthcare facility, it is common for people to witness death. Every event is painful but inscribed in human existence. A nurse usually accompanies a dying individual, and they are obligated to perform their duties in a professional way when handling such a client (Jordan et al., 2022). For nursing staff, similar cases are connected with strong emotions and enormous stress. In 2010, it was discovered that the stress of caring for dying patients was the main cause of high levels of occupational burnout among nurses (Jordan et al., 2022). This means that many nurses, regardless of how they may try to act as if the events are familiar, suffer within themselves.

There are other sources of stress for a nurse while at work, but the death of someone they had become accustomed to appears to be the main one. A study revealed that apart from the shortage of equipment and staff, the client’s demise is usually painful to accept (King, 2022). Additionally, the fear of death is common among healthcare workers and is related to a negative attitude toward caring for a dying client (King, 2022). The passing conflicts with the uncertainty of the therapeutic effect and superiors caused enormous, more than that which is experienced in other areas of work. There is an insufficient finding on the emotions that accompany medical workers while working with a critically sick person (King, 2022). In addition, there are few articles that highlight the methods of dealing with feelings related to dying patients. Helping someone in the last days of their lives is not easy; thus, the great load is associated with challenging circumstances.

The capacity to discover oneself when a client passes, the need to support the deceased’s family, and the necessity to handle individual emotions, are factors that impact the behavior of nurses as professionals and humans. In the present situation of the healthcare system, where there are cases of death due to COVID-19, the nurses are the most stressed (King, 2022). Generally, a patient’s death is regarded as one of the professional circumstances in nursing. Nevertheless, despite having self-control and a calm approach toward the problem, there are still some emotions members of the healthcare need to understand how to handle. The most common consist of anger, sorrow, helplessness, and abandonment.

Family Stress Affecting Work and Vice Versa

During the COVID-19 pandemic, nurses are among the people that were tasked with helping the infected and those in quarantine. It reached an extent they could no longer attend to their private lives since it was feared that they could infect their relatives or loved ones from interaction with patients. For those who were single before the start of the pandemic, it was easier to deal with the times.

However, it became harder for individuals in committed relationships, especially where children were involved. This resulted in stress for both the nurses and their loved ones. As the cases continued to rise, it became even harder for medical experts to claim when the situation would change (King, 2022). The number of deaths rose faster than initially expected, and the fear of losing one’s people increased. This means that, on the one hand, the nurses would think about their families and what would happen if they died. On the other hand, their families, children, spouses, and relatives would have thoughts about what challenges they were facing while at work. This situation impacted both parties negatively, especially emotionally (Kyron et al., 2019). Such sacrifices were noted by the government and citizens too, who considered the healthcare workers heroes of the time (King, 2022). In spite of the importance of family, the nurses understood that they had a role to play in the fight against the spreading of the pandemic and the death of patients.

Recommendation

Similar to how exercise is recommended to help other workers and people, in general, deal with stress, it can aid nurses in coping with their situations. Therefore, organizations should introduce mandatory physical training three or more times every week for nurses (Lowe et al., 2022). This has been proven to be an effective way of reducing stress. It helps to lower adrenaline and cortisol and boost endorphins. It improves energy, which means that the nurses will not be exhausted at the end of their shifts. Even though they work for long hours, it would be appropriate to create short periods on their schedule to allow exercising. In addition to that, listening to music while physically engaging one’s body is helpful in lowering stress levels.

Apart from exercising, it is recommended that nurses find people within the nursing field to talk to about stress. Organizations can create support groups for the employees where they get to meet therapists and other professionals qualified to educate them on an issue such as stress (Trépanier et al., 2022). After long shifts, it has been discovered to be helpful to talk to someone concerning the emotional situation in which a person. A therapist, especially if the stressor involves grief after witnessing the death of a patient, maybe the best. In addition to that, there are fellow nurses who undergo the same issues. Sharing and discovering that others are experiencing similar problems will allow most to understand that they are not alone. This will release the emotional burden that exists on the healthcare workers.

The paper has addressed the stress factors in the workplace and programs organizations could implement to help employee stress. For instance, it has been noted that toxic culture, bullying and harassment, relationships at work, and the death of patients and family are among the major stress factors for nurses. Stress has been identified as how the human body reacts to a change, while a stress factor or stressor is a situation that causes the experience. It is essential to understand that toxicity in the workplace environment can result in the demoralization of employees. This is closely related to bullying and harassment initiated by the more experienced workers over new and younger nurses.

The impact of low morale among nurses, similar to any other employee in other sectors, is low productivity. Since the role of a nurse is caring for patients seeking medical attention, it becomes harder for the latter to get better as the caregiver’s emotional state is affected. It is the responsibility of leaders in the nursing field to ensure that their employees are protected from such situations. Apart from harming the quality of services of a particular hospital, it could impact the whole industry. Adopting programs, including workout sessions and support groups for healthcare workers, can be effective in guaranteeing a healthy mental space for the employees.

Chachula, K. M., & Ahmad, N. (2022). Professional quality of life, stress, and trauma in nursing students: Before and during the novel coronavirus pandemic . Psychological Trauma: Theory, Research, Practice, and Policy . Web.

Díaz McConnell, E., Sheehan, C. M., & Lopez, A. (2022). Intersectional and social determinants of health framework for understanding Latinx psychological distress in 2020: Disentangling the effects of immigration policy and practices, the Trump Administration, and COVID-19-specific factors . Journal of Latinx Psychology . Web.

Jordan, M. M., Freytes, I. M., Orozco, T., Tuan, A. W., Dang, S., Rutter, T., & Uphold, C. R. (2022). The RESCUE problem-solving intervention for stroke caregivers: A mixed-methods pilot study . Rehabilitation Psychology . Web.

King, E. A. (2022). Work-related trauma exposure: Influence on child welfare workers’ mental health and commitment to the field . Traumatology . Web.

Kyron, M. J., Rikkers, W., LaMontagne, A., Bartlett, J., & Lawrence, D. (2019). Work-related and nonwork stressors, PTSD, and psychological distress: Prevalence and attributable burden among Australian police and emergency services employees . Psychological Trauma: Theory, Research, Practice, and Policy . Web.

Lowe, S. R., James, P., Arcaya, M. C., Vale, M. D., Rhodes, J. E., Rich-Edwards, J., Roberts, A. L., & Koenen, K. C. (2022). Do levels of posttraumatic growth vary by the type of traumatic event experienced? An analysis of the Nurses’ Health Study II . Psychological Trauma: Theory, Research, Practice, and Policy , 14 (7), 1221–1229. Web.

Trépanier, S.-G., Peterson, C., Ménard, J., & Notelaers, G. (2022). When does exposure to daily negative acts frustrate employees’ psychological needs? A within-person approach . Journal of Occupational Health Psychology . Web.

  • Chicago (A-D)
  • Chicago (N-B)

IvyPanda. (2024, April 13). Stress Factors in the Workplace. https://ivypanda.com/essays/stress-factors-in-the-workplace/

"Stress Factors in the Workplace." IvyPanda , 13 Apr. 2024, ivypanda.com/essays/stress-factors-in-the-workplace/.

IvyPanda . (2024) 'Stress Factors in the Workplace'. 13 April.

IvyPanda . 2024. "Stress Factors in the Workplace." April 13, 2024. https://ivypanda.com/essays/stress-factors-in-the-workplace/.

1. IvyPanda . "Stress Factors in the Workplace." April 13, 2024. https://ivypanda.com/essays/stress-factors-in-the-workplace/.

Bibliography

IvyPanda . "Stress Factors in the Workplace." April 13, 2024. https://ivypanda.com/essays/stress-factors-in-the-workplace/.

  • Bullying and Harassment in the Healthcare Workplace
  • Bullying in the Workplace Old Nurse to New Nurse
  • The Problem of Bullying
  • Bullying of Nurses During the COVID-19 Pandemic
  • Bullying in the Workplace as a Psychological Harassment
  • Problem of the Managing Bullying and Harassment in the Workplace
  • Bullying and Worker’s Harassment in Western Australia
  • Bullying in Nursing: Causes and Outcomes
  • Power Harassment: Sexual Harassment in the Workplace
  • Bullying in School
  • Organ Donations: Cause and Effect
  • Discussion: Epigenetics and Hypertension
  • Workplace Incidences Prevention
  • Discussion: Legalizing Drugs of Abuse
  • Increase in Mental Health and Depression Rates: An Overview

Read our research on: Gun Policy | International Conflict | Election 2024

Regions & Countries

Political typology quiz.

Notice: Beginning April 18th community groups will be temporarily unavailable for extended maintenance. Thank you for your understanding and cooperation.

Where do you fit in the political typology?

Are you a faith and flag conservative progressive left or somewhere in between.

research paper on stress in the workplace

Take our quiz to find out which one of our nine political typology groups is your best match, compared with a nationally representative survey of more than 10,000 U.S. adults by Pew Research Center. You may find some of these questions are difficult to answer. That’s OK. In those cases, pick the answer that comes closest to your view, even if it isn’t exactly right.

About Pew Research Center Pew Research Center is a nonpartisan fact tank that informs the public about the issues, attitudes and trends shaping the world. It conducts public opinion polling, demographic research, media content analysis and other empirical social science research. Pew Research Center does not take policy positions. It is a subsidiary of The Pew Charitable Trusts .

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
  • Int J Environ Res Public Health

Logo of ijerph

Occupational Stress and Employees Complete Mental Health: A Cross-Cultural Empirical Study

Alcides moreno fortes.

1 School of Psychology, South China Normal University, Guangzhou 510631, China; moc.liamtoh@setrofotilez

2 Center for Studies of Psychological Application, South China Normal University, Guangzhou 510631, China

3 Guangdong Key Laboratory of Mental Health and Cognitive Science, South China Normal University, Guangzhou 510631, China

4 Key Laboratory of Brain, Cognition and Education Sciences, South China Normal University, Ministry of Education, Guangzhou 510631, China

E. Scott Huebner

5 Department of Psychology, University of South Carolina, Columbia, SC 29208, USA; ude.cs.xobliam@renbeuh

Given the shortcomings of previous research on occupational stress and mental health (e.g., predominantly in Western, educated, industrialized, rich and democratic (WEIRD) societies, based on the traditional mental health model and a lack of comparative studies), this study aimed to (a) examine the relationship between occupational stress and complete mental health among employees in Cabo Verde and China, and also explored the mediation and moderation roles of burnout and optimism in accounting for the empirical link. Mental health was defined as comprised of two distinguishable factors: positive and negative mental health. The Pearson correlation test, structural equation modeling (SEM) analysis, bootstrap analysis, hierarchical moderated regression and an independent t-test were used to analyze the data. The results indicated that, in both countries, occupational stress showed a negative relation to positive mental health and lower psychopathology symptoms—and job burnout mediated the relation between occupational stress and mental health. Optimism moderated the relation between occupational stress and burnout, but not the relation between occupational stress and complete mental health. The results are interpreted in light of the comparative framework.

1. Introduction

Occupational stress is generally acknowledged as a global phenomenon with significant health and economic consequences in both developed and developing countries [ 1 ]. Generally defined as a gradual process in which individual cognitive assessments of occupational stressors generate adverse health with severe behavioral consequences [ 2 ], occupational stress results from a “toxic” work environment such as poor control, high work demands, lack of information [ 3 ], extreme pressure [ 4 ] and low decision-making latitude [ 5 ]. Therefore, an employees’ workplace environment is influenced through several organizational resources, including the psychosocial safety climate (PSC). The PSC expresses the preference given to psychological health and well-being in the workplace [ 6 ], i.e., it is an effort made by management to promote an environment conducive to employees’ health, using policies, practices and procedures aimed at protecting and improving the health and psychological security of the employee [ 7 ]. Thus, workplaces with poor PSCs are prone to occupational stress [ 6 ].

In recent decades, occupational stress represents a large, complex and costly phenomenon in the workplace worldwide [ 8 ]. According to the International Labor Organization (ILO) [ 1 ], the workplace environment has been severely affected by globalization and the global financial crisis, leading to an increase in demand as well as stress and related problems. For instance, in the U.K., stress is the leading cause of absenteeism at work [ 9 ]. In the European Region, around half of the workers found stress ‘commonplace.’ [ 10 ]. Concerning the U.S., 83% of employees experience occupational stress [ 11 ]. Alongside the impact on employees’ health and well-being, the effect of occupational stress on the economy is quite notable [ 12 ]. For example, problems arising from stress cost U.S. enterprises approximately USD 300 billion in health-care [ 11 ]. As the ILO [ 1 ] reported, the estimated employers’ total annual cost with mental health disorders among their staff was nearly £26 billion in the U.K., and €617 billion in Europe.

As illustrated above, occupational stress is deemed as a harmful part of the workplace environment [ 13 ], which may severely compromise employees’ well-being, thereby provoking health-related impairments globally [ 14 ]. Thus, this issue has driven the World Health Organization (WHO) to acknowledge the importance of mental health prevention and promotion at the workplace worldwide. Consequently, a growing number of companies, scientific resources, as well as business educators and practitioners have started to devote attention to the employees’ psychological health [ 13 , 14 ]. Furthermore, a significant number of studies has been developed over the last forty years in several occupations [ 9 ], e.g., community health-care [ 15 ], police officers [ 16 ], firefighters [ 17 ], teachers [ 18 ], manufacturing workers [ 13 ] and correctional officers [ 19 ] across different countries. Aimed at helping professionals to comprehend its causes, the key relationships with outcomes that are essential to employees as well as organizational functioning and strategies to mitigate its pervasiveness [ 9 , 14 ].

However, most research in this field has been carried out in Western, educated, industrialized, rich and democratic (WEIRD) societies [ 20 ]. Furthermore, while occupational stress has been at the vanguard of organizational research for many decades [ 14 ], little attention has been devoted to cross-cultural studies [ 20 ]. As Kahn and Byosiere [ 2 ] outlined, cross-cultural differences add immense complexity and difference to the work stress process and understanding these differences, as well as the psychological processes that sustain them, are becoming increasingly necessary [ 14 ]. Thus, the present study aimed to examine cross-cultural differences in the relationship between occupational stress and complete mental health among African (Cabo Verde) and Asian (China) employees and also explored the mediation and moderation effects of burnout and optimism in accounting for the relations.

The suggested theoretical framework described in Figure 1 will support the present study. Specifically, occupational stress is linked to positive and negative mental health, mediated by burnout and moderated by optimism. In this study, two samples of Cabo Verdean and Chinese employees were tested respectively, and the results are explained according to the culture of the associated country.

An external file that holds a picture, illustration, etc.
Object name is ijerph-17-03629-g001.jpg

The theoretical framework. Note. OSQ—GV (Occupational Stress Questionnaire—General Version); MBI-GS (Maslach Burnout General Inventory Scale including exhaustion, cynicism, and professional efficacy); LOT-R (Revised Life Orientation Test), MHC—SF (Mental Health Continuum—Short Form including psychological, emotional and social well-being); MHI-5 (Five-item Mental Health Inventory including anxiety and depression).

1.1. Dual-Factor Model of Mental Health

The conception of mental health has been historically based on the presence or absence of symptoms of psychopathology [ 21 ]. Over the past two decades, emerging research in mental health [ 22 ] has been highlighting the importance of well-being as an essential indicator of mental health [ 23 ]. As the WHO [ 24 ] observed, mental health is more than the absence of mental disorders or disabilities, but also a complete state of well-being in which the subject uses their potential for the benefits of the community, involving the ability to maintain meaningful relationships with others [ 25 ]. In order to address mental health comprehensively, the dual-factor model [ 22 ] posits mental health as a complete state that integrates both the absence of illness and a high level of subjective well-being. Moreover, this model incorporates two related but distinct dimensions: negative (e.g., symptoms of psychopathology) and positive (e.g., subjective well-being) mental health [ 26 , 27 ].

1.2. Occupational Stress and Mental Health

Stress is an integral part of employees’ life and happens in a wide variety of job circumstances [ 3 ]; the long-term excessive stress could lead to psychological problems like depression and anxiety [ 1 ]. Therefore, with the rapid development and socio-economic transformation in recent decades, the study of mental health based on the relationship with occupational stress has gradually become an expanding area of research [ 18 , 28 ]. As a consequence, there has been significant research interest in attempting to explain the link between stress and mental health. Empirical studies demonstrated that occupational stress is a significant predictor of anxiety [ 29 , 30 , 31 ]. Furthermore, it inversely correlated with psychological well-being [ 32 ] and positively associated with depressive symptoms [ 13 , 33 ]. Nakao [ 34 ] recognized that occupational stress is a meaningful cause for mental health all over the world. It thus seems reasonable to expect a link between occupational stress and complete mental health. Hence, we hypothesized that:

There is a negative relationship between occupational stress and complete mental health (positive and negative) among employees.

1.3. Role of Burnout

Occupational stress is the most significant cause of burnout [ 1 ], which can also indirectly affect mental health [ 15 ]. To explore this phenomenon, we need to examine the mediation effect of burnout in the relationship between occupational stress and complete mental health. Burnout is a syndrome emerging from chronic workplace stress that has been poorly administrated, and it is distinguished by the following three dimensions: emotional exhaustion, cynicism associated with someone’s work, and reduced professional efficacy [ 35 ]. Following this, it was noted that stress and burnout are closely related variables. Wang et al. [ 36 ] demonstrated that a high level of occupational stress is linked to more burnout symptoms, and Hayes and Weathington, [ 37 ] and Wang Ziyu et al. [ 13 ] found that occupational stress is positively correlated with burnout. More recently, Chen et al. [ 38 ] also found that burnout mediated the relation between occupational stress, depression and anxiety symptoms among young nurses. Increased burnout and exhaustion itself might lead to different adverse outcomes [ 39 ]. Based on the extant literature, we thus hypothesized that:

Employee’s occupational stress is positively associated with burnout.

Burnout is negatively associated with positive and negative mental health

Burnout will mediate the relationship between occupational stress and complete mental health (positive and negative mental health)

1.4. Roles of Optimism

Researchers have defined optimism as the generalized expectancy for positive outcomes [ 40 ]. It is a tendency to expect better events in the future, a tendency to anticipate pleasant results in life [ 41 ]. Caprara et al. [ 42 ] observed that optimism is an important personal characteristic that contributes to the improvement of good mental health; i.e., optimistic people have better abilities to deal with life obstacles. Furthermore, it has a significant impact on an individual’s stress-coping experience [ 43 ] and can be helpful to improve mental health [ 44 ]. Large-scale studies corroborated findings that higher optimism and stress are negatively associated [ 37 ]. Moreover, optimism or positive expectation about future events can also play a significant influence to better people’s well-being and stress-coping mechanisms [ 45 , 46 ].

In the occupational stress research field, the Job Demands-Resources Model (JD–R model) [ 47 ] postulated that the characteristics of the individual could act as moderators in the adverse impact of stress on the employee’s mental health. Moreover, researchers have mostly focused on the association between occupational stress and psychopathology symptoms, paying little attention to optimism as a moderating variable [ 33 ]. For example, Chang [ 48 ] observed a significantly moderated effect of optimism in the occupational stress–psychological well-being relationship, and Banerjee [ 33 ] and Romswinkel et al. [ 49 ] found that optimism acts as a buffer in the relationship between occupational stress and depressive symptoms. Rollei and Savicki [ 50 ] also found the moderated effect of optimism in the chronic stress and burnout relationship. Based on the above findings, we therefore further hypothesized that optimism would moderate the impact of occupational stress on employees’ burnout and complete mental health. Hence, we proposed:

Optimism will moderate the relationship predicted in hypothesis H2 (occupational stress–burnout).

Optimism will moderate the relationship predicted in hypothesis H1 (occupational stress–mental health).

Previous research [ 51 , 52 ] showed that Western individuals usually have higher optimism scores compared with Eastern individuals. Thus, we hypothesized that:

Cabo Verdean employees will report higher optimism scores compared to Chinese employees.

1.5. Cultural Context and Stress Management

According to Pasca and Wagner [ 53 ], employees’ cultural variability is a fundamental element in understanding the complex phenomenon of workplace stress. Thus, existing research on stress and mental health has been focused on European and South American societies [ 54 ], especially without adequate representation from African samples [ 55 ]. Although Africa will make up a quarter of the world’s population by 2050 [ 56 ], and that China is twenty percent of the world’s population [ 57 ].

Moreover, it has been called into question whether knowledge of this area would apply to non-European or American cultures. For example, Burke [ 20 ] questioned the extent to which research findings from advanced industrialized countries applied to the developing countries. Yew et al. [ 58 ] argued that perceiving and the management of stress could be affected by the cultural environment to which the person belongs. Aldwin [ 59 ] agreed that stress management was profoundly affected by cultural context, and in the same line of thought, Wahid Ahmad Dar [ 60 ] considered that there are culturally specific behaviors in the stress management process. Triandis and Suh [ 61 ] concurred that the work attitude and managerial behavior of people in collectivist and individualist cultures are different. Similarly, the theory of culture’s consequences [ 62 ] questioned whether the occupational stress process manifests in a different way according to the countries [ 14 ].

In summary, substantial evidence has demonstrated that cultural context might be influential in stress management. Hence, cross-comparison studies on stress and mental health between African and Asian countries are needed. The prevalence and strength of studies involving different contexts allow testing theories and hypotheses about the differences in the functioning of psychological mechanisms, as well as to discover reasons why the differences exist [ 20 , 63 ].

1.6. Current Study

This study aimed at examining the cross-cultural differences in the relationship between occupational stress and complete mental health among Cabo Verdean and Chinese employees, including the examination of the mediating role of burnout and the moderating role of optimism in accounting for the empirical link. Therefore, the findings are explained in light of the comparative framework.

2. Materials and Methods

2.1. participants.

The data was collected from 440 employees in Cabo Verde and China. The demographic data included gender, age, marital status, school background, work year and weekly work time. Marital status was categorized by married, single and divorced. School background was distinguished by a high school degree and a university degree. Weekly work-time was defined as 40 h and more than 40 h.

In Cabo Verde, the participants were 263 employees with ages ranging from 20 to 60 years old, and the majority were in the range 30–39 years old. Over half of the sample was male (male = 71.5%, female = 28.5%), and they were recruited from 9 islands of the Cabo Verde archipelago. In China, the participants were 177 employees, with ages ranging from 20 to 60 years old and the majority were in the range 21–29 years old. Over half of the sample was female (female = 56.5%, male = 43.5%), and they were recruited from China. Additional information about participant characteristics can be found in Table 1 .

Demographic profile of the overall sample.

2.2. Measures

2.2.1. occupational stress.

Occupational stress was assessed through the Occupational Stress Questionnaire—General Version (OSQ—GV) [ 64 ], consisting of two distinct parts. First, the general level of stress was assessed through a single item (0 = no stress, 2 = moderate stress and 4 = high stress). In the second section, there were 24 items related to potential sources associated with professional activity. The questions were distributed in seven subscales (relationship with clients, relationship with bosses, relationship with colleagues, overwork, career and remuneration, family problems and working conditions), answered on a five-point Likert scale (0 = no stress, 2 = moderate stress and 4 = high stress). The score was the result of adding the items of each dimension and subtracting the values found by the total number of items in the subscale. Therefore, higher values reflected a higher perception of stress in each of the domains evaluated. In this study, Cronbach’s alpha found excellent reliability α = 0.93 for Cabo Verdean and α = 0.94 for Chinese.

2.2.2. Positive Mental Health

Positive mental health was assessed through the Mental Health Continuum—Short Form (MHC—SF) [ 26 ], which was composed of 14 items on a six-point Likert scale (1 = never, to 6 = every day). The MHC—SF was subdivided into three subscales that aimed to evaluate psychological, emotional and social well-being, with a reliability test above 0.80 for each subscale and on the overall scale [ 26 ]. In the present study, Cronbach’s alpha found good reliability α = 0.89 for Cabo Verdean and α = 0.94 for Chinese.

2.2.3. Negative Mental Health

The psychopathology symptoms were assessed through the Five Item Mental Health Inventory (MHI-5) [ 65 ], which was an integral part of the eight independent subscales that make up the SF-36 health-related quality of life measure (SF-36 HRQOL) [ 66 ]. MHI-5 was comprised of five questions to assess symptoms of depression (item b, d and e) and anxiety (a and c) in clinical or non-clinical groups, on a 5-point Linkert scale (1 = all, two = most, 3 = some, 4 = a little, or 5=none of the time). The scores ranged from 0 to 100, and the lower scores expressed more severe depression and anxiety symptoms, while high scores signified the absence of symptoms. In this study, Cronbach’s alpha found good reliability α = 0.79 for Cabo Verdean and α = 0.79 for Chinese.

2.2.4. Burnout

Burnout was assessed through the Maslach Burnout General Inventory Scale (MBI-GS) [ 67 ], including 16 items on a 7-point Likert scale (0 = never, to 6 = daily). MBI-GS was divided into three subscales that assessed emotional exhaustion, cynicism and also professional effectiveness. Burnout was indicated by high scores of exhaustion, cynicism and lower scores of effectiveness. In this study, Cronbach’s alpha found good reliability α = 0.72 for Cabo Verdean and α = 0.87 for Chinese.

2.2.5. Optimism

Optimism was assessed through the Revised Life Orientation Test (LOT-R) [ 68 ], which aims to assess optimism and pessimism (life-orientation), including only ten items (optimism = 3 items, pessimism = 3 items, and distractors = four items). The items are on a 5-point Likert scale (e.g., (a) = agree a lot (b) = agree a little, (c) = neither agree nor disagree, (d) = disagree a little and (e) = disagree a lot). The level of optimism was achieved through the general taming of the items. Internal consistency was 0.70, the test-retest was 0.68 to 0.79 [ 68 ]. In this study, the Cronbach’s alpha coefficient found optimism α = 0.86, pessimism α = 0.67 for Cabo Verdean and optimism α = 0.75, pessimism α = 0.67 for Chinese.

2.3. Procedure

The ethical approval of the present research was provided by the School of Psychology Research Ethics Committee, South China Normal University. All procedures performed in this study involving human participants were following the ethical standards of the institutional or national research committee and with the 1964 Helsinki declaration and its later amendments or comparable ethical standards. A questionnaire was distributed to the employees in Cabo Verde and China. All the measurements used were in the Portuguese version for Cabo Verdean employees and in the Mandarin version for the Chinese. Considering that the original version of the Occupational Stress Questionnaire—General Version (OSQ—GV) [ 64 ] was in Portuguese, it was translated into Mandarin, followed by a retro translation, turning the items back into Portuguese, to ascertain the quality of the first translation. First, the scale was sent to a small group of employees, to investigate any difficulties in the instruction level and the understanding of the items themselves. After this was performed, a semantic validation allowed the building of an online version. The dissemination of the study and the request for collaboration with the employees was made through the social network, e.g., Facebook, Viber, WhatsApp and Messenger in Cabo Verde, and WeChat and QQ in China. The free content term was on the first page and the condition to proceed depended on the acceptance of the search terms. The link to join in the survey was made available for 30 days and participants did not receive any reward for taking part in the research. Assuming the first objective of the study, the internal consistency of all instruments used in this study was analyzed using Cronbach’s alpha.

2.4. Analytic Strategy

The data were collected in an online survey program hosted by Google Docs and Questionnaire Star. Because an answer was mandatory in all questions, there were no-missing answers among the participants. The data were subsequently included in five procedures, namely: first, we used the Statistics SPSS 24.0 software (IBM, Armonk, NY, USA) to statistically analyze the properties of the collected data, including the means, standard deviations and the Pearson correlation analysis. Second, structural equation modeling (SEM) was applied in AMOS 24.0 to examine our two models, in which occupational stress directly predicted the complete mental health and also examined the mediating role of burnout and the moderator role of optimism. The Tucker Lewis index (TLI), the Incremental Fit index (IFI), the comparative fit index (CFI), and the and the root mean-square error of approximation (RMSEA) [ 69 , 70 ] were used to assess the model fit, and the model will be accepted when the CFI, IFI and TLI values are above 0.90 [ 71 ]. The RMSEA below 0.06 was considered a good fit, while values between 0.06 and 0.10 could also be accepted [ 72 , 73 ]. Third, the bootstrap 1000 sample was performed to calculate the estimates of the mediation effects [ 74 ]. Fourth, the regression analyses were carried out to investigate the moderating effect of optimism in the relationship between occupational stress and complete mental health, as well as occupational stress and burnout. It is worth mentioning that the mediation and moderation analyses were conducted separately. Finally, the independent t-test was used to examine the potential statistical differences in the overall measures between the Cabo Verdean and Chinese contexts. Statistical significance was defined as a p -value of <0.05.

3.1. Preliminary Analyses and Descriptive Statistics

Four hundred and forty employees participated in this study. Pearson correlations, structural equation modeling (SEM) analysis, bootstrap analysis, hierarchical moderated regression and the independent t-tests were carried out as main the statistical methods of this study. Table 2 displayed the variable means (M.D.), standard deviations (S.D.), inter-correlations of the main variables and the internal reliabilities (α). The data were normally distributed since the skewness value was <2.0 and the kurtosis value was <7.0 [ 75 , 76 ]. The Cronbach’s alpha coefficients were in the range of 0.67–0.94, showing the satisfactory internal reliability of all the scales. In order to examine the relationships between stress, burnout, optimism and complete mental health within these two samples, Pearson correlation coefficients were computed for every study variable (see Table 2 ).

Cronbach’s alpha coefficients, the means, the standard deviations and the Pearson correlation coefficients.

Note. (** p < 0.01. * p < 0.05.) (Occupational Stress Questionnaire—General Version: OSQ—GV); (Maslach Burnout General Inventory Scale: MBI-GS); (Revised Life Orientation Test: LOT-R) (Mental Health Continuum—Short Form: MHC—SF); (Five-item Mental Health Inventory: MHI-5).

As expected, occupational stress had a significant reverse correlation with positive mental health ( r = −0.30, p < 0.01) and negative mental health for the Cabo Verdean employees ( r = −0.25 p < 0.01). Similarly, occupational stress had a significant reverse correlation with positive mental health ( r = −0.35, p < 0.01) and negative mental health ( r = −0.29, p < 0.01) for the Chinese employees. Additionally, occupational stress and optimism had a significant reverse correlation ( r = −0.26, p < 0.01), in Cabo Verdean employees, and a significant positive correlation ( r = 0.27, p < 0.01) in Chinese employee. The results also indicated that occupational stress and burnout had a positive correlation for both Cabo Verdean ( r = 0.63, p < 0.01) and Chinese employees ( r = 0.62, p < 0.01). Burnout and complete mental health had a reverse correlation for Cabo Verdean ( r = −0.22, p < 0.01; r = −0.30, p < 0.01) and Chinese employees ( r = −0.27, p < 0.01; r = −0.14, p > 0.05). Hence our first H1, second H2 and third H3 hypotheses were supported.

3.2. Structural Equation Modelling Results

A priori, the direct effect of occupational stress (OSQ—GV) on positive (MHC—SF) and negative mental health (MHI-5) without mediators were tested. The directly standardized path coefficient was significant, β = −0.32, p < 0.01 and β = −0.31, p < 0.01 for the Cabo Verdean employees and β = −0.34, p < 0.01 and β = −0.38, p < 0.01 for the Chinese employees. For evaluating the fit of the total model, the SEM technique was employed in AMOS. The model included four latent factors, namely occupational stress, burnout, positive and negative mental health and 14 observed variables for each country. The complete structural equation models of both countries are exposed in Figure 2 .

An external file that holds a picture, illustration, etc.
Object name is ijerph-17-03629-g002.jpg

The Mediation model test results. Note. OSQ—GV = S1 to S7 subscales of occupational stress; MBI-GS = (EX = exhaustion–energy, CY = cynicism–involvement); MHC—SF = (PWB = Psychological well-being, EWB = Emotional well-being, SWB = Social well-being); MHI-5 = (DEP = Depression, ANX = Anxiety), (* p < 0.05.).

Consequently, the generic model, which posited three burnout dimensions as mediators of the effects of occupational stress on complete mental health (OSQ—GV and MHI-5), fit poorly: CFI = 0.92; TLI = 0.90; IFI = 0.92; RMSEA = 0.09; and CFI = 0.89; TLI = 0.86; IFI = 0.89; RMSEA = 0.11 for the Cabo Verdean and Chinese employees, respectively. After the subsequent computing of the correlations of the residual terms, the revised models exhibited an acceptable fit to the data: CFI = 0.93; TLI = 0.91; IFI = 0.93; RMSEA = 0.08, for the Cabo Verdean employees and CFI = 0.94; TLI = 0.92; IFI = 0.94; RMSEA = 0.08 for the Chinese employees. The standardized path coefficients from OSQ—GV to MHC—SF and MHI-5 in both groups were all statistically insignificant (see Figure 2 ).

3.3. Mediation Effect of Burnout on Occupational Stress and Mental Health

The bootstrap process suggested by Shrout and Bolger [ 77 ] was used to examine the statistical significance of the indirect effects in the mediation model. As shown in Figure 2 , the analyses yielded coefficients statistically significant of both the a and b paths, in the predicted way; hence, this provided the support to mediation. Employing bias-corrected bootstrapping with 1000 resamples, the mediation effect of occupational stress on positive mental health was −0.36 with a 95% CI: −0.77 to −0.17, on negative mental health was −0.24 with a 95% CI: from −0.64 to −0.03, for Cabo Verdean employees. Concerning the Chinese employees, this was −0.15 with a 95% CI: from −0.25 to −0.005, and −0.24 with a 95% CI: from −0.51 to −0.05 for positive and negative mental health, respectively.

Additionally, the mediation was supported since the 95% CI did not hold zero, as well as had a statistically significant indirect effect. Table 3 illustrates the bootstrap results for mediation analysis, and Figure 2 illustrates the direct path coefficient of predator (OSQ—GV) to criteria (MHC—SF and MHI-5) which for the Cabo Verdean was β = 0.04 p = 0.71; β = −0.02 p = 0.83 and for the Chinese employees β = −0.17 p = 0.14; β = −0.11 p = 0.47, respectively. Therefore, burnout had a full mediation in the occupational stress and positive and negative mental health relationship for both countries. Hence, these results also endorsed our fourth H4 hypothesis.

Bootstrap results for the indirect effect.

Note. OSQ—GV = occupational stress; MBI-GS = burnout; MHC—SF = positive mental health; MHI-5 = negative mental health relation.

3.4. Moderation Effect of Optimism

Using hierarchical moderated regression, we tested the hypothesized moderating role of optimism on the occupational stress (OSQ—GV), burnout MBI-GS and complete mental health (MHC—SF and MHI-5) relationship. Before testing the analyses, all the terms were centered on reducing multicollinearity among variables, and the interaction variables were added. Furthermore, it was required to determine whether the goodness-of-fit indicators of the Cabo Verdean and Chinese models reached acceptable levels. The fit of the Cabo Verdean model was CFI = 0.99, TLI = 0.97, IFI = 0.99, RMSEA = 0.04. The fit of the Chinese model was CFI = 0.99, TLI = 0.95, IFI = 0.99, RMSEA = 0.06.

The hypothesis H6 suggested that optimism moderated the occupational stress and complete mental health relationship. The findings revealed that the interaction coefficient for occupational stress and optimism was not statistically significant in both groups, as illustrated in Figure A1 (see Appendix A ). Additionally, hypothesis H5 also proposed that optimism moderated the occupational stress and burnout relationship. The findings revealed that the interaction coefficient for occupational stress and optimism were statistically significant (β = −0.17, p < 0.01) for the Cabo Verdean employees, and were also significant (β = 0.14, p < 0.01) for the Chinese employees.

As depicted in Figure 3 , at high levels of occupational stress, employees with high optimism scores displayed lower burnout levels than individuals with low optimism scores. This result suggests that optimism is associated with a reduced risk of job burnout. Stress significantly affects job burnout, i.e., stress and burnout levels increase in the direct sense, and higher levels of optimism contribute to decreased levels of burnout. In short, these results are consistent with the expected relationship; which is when Cabo Verdean employees exhibited high optimism or expectations of positive outcomes, they had fewer chances of experiencing burnout. In the case of the Chinese employees, the results also indicated that optimism had a significant influence on job burnout, which means that at moderate height levels of occupational stress, employees with high optimism scores displayed lower burnout levels than employees with low optimism scores. Thus, the relationship between occupational stress and burnout was a moderator by optimism. Hence, our fifth hypothesis, H5, was supported while our sixth H6, was not supported.

An external file that holds a picture, illustration, etc.
Object name is ijerph-17-03629-g003.jpg

Graphic representation of the interaction between occupational stress and optimism in predicting employee job burnout.

3.5. Differences between Occupational Stress and Complete Mental Health

Table 4 illustrates the M.D. (variable means) S.D. (standard deviations) and the means differences between the two groups of the employees on their occupational stress, burnout, optimism, positive mental health and negative mental health subscales. The independent t-test values detected significant differences in different subscales. Notably, overwork as well as career and remuneration were two significant sources of stress while working conditions was a lesser source of stress for both countries. Moreover, the work overload appeared much healthier ( t = 21.6, p < 0.01) for the Cabo Verdean employees than for Chinese employees. In a study of burnout among Canadian and Chinese employees, Jamal [ 78 ] also found that work overload was much stronger for the Canadian than the Chinese employees. The Chinese employees showed a higher level of well-being than the Cabo Verdean employees, although they had slightly higher symptoms of depression and were less optimistic. Cabo Verdean employees revealed a lower level of stress at work ( t = −11.8, p < 0.01), positive mental health ( t = −12.3, p < 0.01) and higher anxiety and depression symptoms (i.e., lower scores expressed high depression and anxiety symptoms) ( t = −3.82, p < 0.01) than the Chinese.

Mean differences between the Cabo Verdean ( n = 263) and the Chinese employees ( n = 177).

Note. ** p < 0.01. (Occupational Stress Questionnaire—General Version: OSQ—GV); (Maslach Burnout General Inventory Scale: MBI-GS); (Revised Life Orientation Test: LOT-R) (Mental Health Continuum—Short Form: MHC—SF); (Five-item Mental Health Inventory: MHI-5).

As far as the level of burnout, there was a notable difference between the means of the two countries’ employees. Cabo Verdean employees also revealed a lower level of burnout ( t = −20, p < 0.01) than the Chinese employees. According to the burnout cut-off points [ 35 ], the results from both countries showed average levels of exhaustion–energy and cynicism–involvement, and a high level of professional efficacy. Cabo Verdean employees reported a higher level of optimism ( t = 40.4, p < 0.01) compared with the Chinese employees. Therefore, the hypothesis H7 was supported.

4. Discussion

Research on occupational stress can offer an essential function in the development of strategies aimed at the prevention and promotion of mental health at work [ 79 ]. Thus, this study was the first to examine the relationship between occupational stress and complete mental health and the psychological mechanisms that account for such a relationship in two different cultures (Asian and African).

4.1. The Relationship Between Occupational Stress and Complete Mental Health

Consistently with previous studies [ 23 , 80 ], occupational stress had a significant reverse correlation with positive mental health as measured by the MHC—SF, (i.e., psychological, emotional and social well-being) [ 29 ] and was also negatively related to the lower negative mental health measured by MHI-5 (i.e., anxiety and depression symptoms) [ 65 ] in both groups. According to the present finding, employees with higher levels of mental health were more likely to experiences low levels of depression and anxiety symptoms [ 81 , 82 ]. Thus, it supported the assumption that occupational stress represents a risk factor concerning an employees’ mental health. The results also demonstrated a direct and statistically significant correlation between occupational stress and burnout in both cases. Thus, this finding was consistent with previous research [ 13 , 36 , 37 ]. Job burnout and stress increased in the direct sense that may lead to chronic stress, as has been supported in the scientific literature. A significant inverse association between optimism and occupational stress for Cabo Verdean employees and a direct relation for a Chinese employee was also observed. Thus, these correlations were in the expected direction since Eastern individuals usually score lower levels of optimism compared to Western individuals [ 51 , 52 ].

4.2. The Mediating and Moderating Roles of Burnout and Optimism

Concerning the indirect effect of occupational stress via burnout on complete mental health, the results showed that burnout was linked to higher levels of occupational stress, depression and anxiety symptoms, thus playing a significant mediating role in the relationship mentioned above among employees of both countries. Thus, occupational stress can not only directly affect employees’ mental health, but it can also indirectly affect their mental health by increasing employees’ job burnout in both countries. These findings were compatible with the results of antecedent studies [ 39 , 40 ]. It is noteworthy that, according to our results, hypothesis H4 was fully supported in both countries.

Figure A1 depicts the relationship between occupational stress, optimism, burnout and complete mental health (positive and negative mental health). Optimism revealed a similar relationship with stress and burnout in both groups. Therefore, optimism moderated the relationship between occupational stress and burnout in both contexts, so it seems to protect employees from the risk of job burnout. The results were consistent with previous studies [ 50 ]. Thus, positive expectations about the future could prevent the development of work stress by relieving job burnout. It may be that enhancing positive personal characteristics such as optimism could be taken into consideration in improving intervention plans for mental health promotion with Chinese and Cabo Verdean employees. Although there is evidence supporting the moderating roles of optimism in stress and mental health relationships (well-being and symptoms of psychopathology) [ 33 , 48 , 49 ], in the present study optimism did not play a moderating role in the occupational stress and mental health (positive/negative) relationship, in employees from both countries.

4.3. The Differences between Cabo Verdean and Chinese Employees

Chinese employees revealed more occupational stress than Cabo Verdean employees because of Chinas’ rapid economic development, which increases the demands of customers and employers [ 82 ]. The overwork, career and salary are the significant sources of stress in the employees of both countries. Cabo Verdean employees, compared with Chinese employees, reported higher scores of optimism our last hypothesis (H7) predicted. This result was consistent with Chang, [ 51 ] and Lai and Yue [ 52 ], studies that recognized that Western individuals, in general, report higher optimism scores than Eastern individuals.

However, both groups of employees displayed high levels of positive mental health but compared to the standards of the three dimensions of positive mental health (psychological, emotional and social well-being), the Chinese employees revealed higher levels compared to the Cabo Verdean employees. It shows that in the case of Chinese employees, there is an influential culture of cooperation (collectivism) that facilitates support among them, which helps them to cope with negative feelings in a better way than employees in western countries (individualism) [ 52 , 78 ]. The Chinese people belong to a collectivist culture, characterized by a strong sense of belonging and mutual help throughout life [ 59 ], where harmony should always be preserved. Thus, it always promotes social, emotional and psychological well-being in Chinese people.

4.4. Strengths, Limitations and Future Directions

The current study investigated occupational stress in samples of Cabo Verdean and Chinese employees. As indicated, most studies in this field have focused on western industrialized countries, ignoring Asian and African samples. This study examined occupational stress, taking into consideration the cultural differences. Moreover, mental health has usually been studied based on traditional models, while in our case, mental health was considered as a whole as advised by Greenspoon and Saklofske [ 22 ]. Thus, our findings demonstrated that the constructs of stress, burnout, optimism as well as positive and negative mental health could provide a meaningful basis for understanding the risk and protective factors of mental illness in the workplace environment in different cultural contexts. Therefore, by following the mechanism of this relationship, healthcare professionals and employers can develop an intervention system comprehensively in the workplace. Thus, companies can develop a more adapted intervention in their programs to establish strengths (optimism) for coping with stress at work to promote mental health in the workplace. Increased optimism might be promising in decreasing burnout among employees with high occupational stress in both countries, since the success of any organization depends upon the well-being of its collaborators.

The study presented some shortcomings that deserve to be pointed out. This study was conducted on a relatively small cohort of Cabo Verdean and Chinese samples. Moreover, this kind of design usually may not be possible to make a solid conclusion about causality. The participants both in Cabo Verde and China completed a large (75 items) online questionnaire without the presence of the researchers, which could lead to a lack of honesty in some answers by some participants. Although the samples of this study were relatively small, the results were sufficiently robust to suggest that further studies with cultural comparisons would be justified.

5. Conclusions

Our results demonstrated that occupational stress is a risk factor for mental health in employees in both countries. Moreover, it also indirectly impacted positive and negative mental health through burnout. Based on the assumption of positive psychology, personal characteristics such as optimism, can play a fundamental role in buffering the impact of stress on mental health. However, although it failed to moderate the relationship between occupational stress and mental health, in the current study, it did so through the relationship between stress and burnout. This means that positive expectancies about life may be more predisposed to evaluate the pressure as a flexible threat, thereby checking their psychological adjustment to the exhaustion. The significant differences between the employees from both countries in terms of the impact or management of stress proved to be quite substantial, due to the cultural characteristics of each people. Although the Chinese employees were under immense pressure, they revealed better levels of social, emotional and psychological well-being compared to the Cabo Verdean employees, given the influential collectivist culture of the former. However, Asian people cope better with life and control their emotions more effectively because in the collectivist culture, people tend to adjust to the others’ demands [ 83 ], which benefits in the coping process. Thus, cultural aspects and personal characteristics (optimism) should be taken into consideration in mental health protection and promotion at the workplace.

Acknowledgments

We cordially thank Dr. Rui Gomes, from the School of Psychology at the University of Minho, author of the Questionário de Stress Ocupacional—Versão Geral (Occupational Stress Questionnaire—General Version) (QSO—VG), for providing and authorizing its use in our study. This extends to colleagues who contributed to the translation of the material into the Chinese language to be applied to the Chinese employees. Finally, we extend our gratitude to everyone who contributed to this study.

Appendix A. Moderation Model of the Test Results

An external file that holds a picture, illustration, etc.
Object name is ijerph-17-03629-g0A1.jpg

The moderation model of the test results. Note. ZOSQ—GV = occupational stress; ZLOT-R = optimism; ZMBI-GS = burnout; ZMHC—SF = positive mental health, ZMHI-5 = negative mental health relation (Z = means all variables are centred).

Author Contributions

All authors contributed equally to this research. Formal analysis, A.M.F., project administration and supervision, L.T., and E.S.H. All authors have read and agreed to the published version of the manuscript.

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

IMAGES

  1. 🎉 Research paper on stress in the workplace. Managing Stress in the

    research paper on stress in the workplace

  2. The 2020 UK workplace stress survey

    research paper on stress in the workplace

  3. (DOC) Research Paper About Stress

    research paper on stress in the workplace

  4. Full Stress Management Research Paper

    research paper on stress in the workplace

  5. (PDF) IMPACT OF STRESS ON WORK PERFORMANCE OF EMPLOYEES

    research paper on stress in the workplace

  6. 5 Simple Stress Management Strategies for the Workplace [Infographic]

    research paper on stress in the workplace

VIDEO

  1. Paper stress reliever. Kağıtdan stress giderici😉#shorts

  2. How to make paper stress reliever

  3. HOW TO MANAGE STRESS AT WORK

  4. How to make paper stress reliever

  5. Workplace Stress Solutions: Family Support

  6. How stress can improve performance (and also hinder it)

COMMENTS

  1. (PDF) Stress at the Workplace and Its Impacts on Productivity: A

    In every fast-paced surrounding, stress is present in every life aspect, including at the workplace. It is a deeply person-al experience, with various stressors affecting every individual differently.

  2. Workplace stress: A neglected aspect of mental health wellbeing

    Guidelines to improve workplace culture and reduce stress. The World Health Organization has outlined key factors related to stress at workplace and advocated guidelines to redeem them 11.Some factors that cause increased stress at workplace include 'workload (both excessive and insufficient work), lack of participation and control in the workplace, monotonous or unpleasant tasks, role ...

  3. Perceptions of work stress causes and effective interventions in

    The conceptualisation of work stress is of crucial importance when developing interventions for the workplace. Work-related stress is defined as 'a harmful reaction that people have to undue pressures and demands placed on them at work'. 1 As many as 440 000 people in the UK complain of work-related stress, depression or anxiety that makes them ill; nearly 9.9 million work days were lost ...

  4. Work environment risk factors causing day-to-day stress in occupational

    Background While chronic workplace stress is known to be associated with health-related outcomes like mental and cardiovascular diseases, research about day-to-day occupational stress is limited. This systematic review includes studies assessing stress exposures as work environment risk factors and stress outcomes, measured via self-perceived questionnaires and physiological stress detection ...

  5. Work stress, mental health, and employee performance

    Work stress and employee performance. From a psychological perspective, work stress influences employees' psychological states, which, in turn, affects their effort levels at work (Lu, 1997; Richardson and Rothstein, 2008; Lai et al., 2022 ). Employee performance is the result of the individual's efforts at work (Robbins, 2005) and thus is ...

  6. Perceived Stress, Work-Related Burnout, and Working From Home Before

    While social and traditional media discussed work-related stress and burnout during the COVID-19 pandemic, there was little empirical research to examine the phenomena except for a few high-level surveys (Brynjolfsson et al., 2020; Center for National Health Statistics, 2020; CVS Health, 2020; Petterson et al., 2020).Although the quickly forced shift to working from home was brought on by ...

  7. Work and stress: A research overview.

    Stress is a leading cause of ill health in the workplace. This shortform book analyses, summarizes and contextualizes research around stress at work. The book begins by exploring the impact and challenges of technology and the challenging and changing contours and boundaries of the nature of work. Using a behaviour lens, the authors draw on cyberpsychology to illuminate the choices we make to ...

  8. Stress at work

    Work overload (working after hours) or insufficient workload (nothing to do) Discrimination, humiliation, violence, bullying, and harassment at work. Cases of work related stress in the same company. Company situation in terms of finances, organisational changes, and employee turnover. Job insecurity, temporary employment status.

  9. Work, Stress, Coping, and Stress Management

    Work stress is a generic term that refers to work-related stimuli (aka job stressors) that may lead to physical, behavioral, or psychological consequences (i.e., strains) that affect both the health and well-being of the employee and the organization. Not all stressors lead to strains, but all strains are a result of stressors, actual or perceived.

  10. Psychological Resources and Strategies to Cope with Stress at Work

    The connection between hope and stress in the workplace has received little if any attention in research, but there is compelling evidence from hope research in other contexts (e.g., clinical and sports psychology) which suggests that hope may be a positive resource in stressful situations (Avey, Luthans, & Jensen, 2009). Hope may have an ...

  11. Workplace stress and health

    The remainder of the paper is structured as follows. First, some theoretical contributions and previous research in the relevant areas, workplace health, stress, workplace bullying and quality management are briefly overviewed. Next, the methodology of the study is described. Subsequently, the findings are presented, elaborated and discussed.

  12. PDF The Effect of Occupational Stress and Coping Strategies on Mental ...

    academics with the aim of reducing work stress might be seen as a potentially beneficial adaption to difficulties in the modern environments. An opinion supported by Müller and Schumann (2011). However, there is a lack of research on the effect of academics' coping styles on their health and emotional well-being.

  13. Burnout and stress are everywhere

    As in 2020, American workers across the board saw heightened rates of burnout in 2021, and according to APA's 2021 Work and Well-being Survey of 1,501 U.S. adult workers, 79% of employees had experienced work-related stress in the month before the survey.

  14. (PDF) Stress at the Workplace and Its Impacts on Productivity: A

    According to Leung et al. (2010), stress may express nervousness or frustration and could affect performance and mental health as well. Bowen et al. (2013) also found similar results in their research. 2.2 Stress at the Workplace Work stress is dependent on the work itself.

  15. Organizational Best Practices Supporting Mental Health in the Workplace

    Workplace stress has been identified as one of the biggest sources of employee anxiety and depression, and chronic stress can even have prolonged negative impacts on one's psychological and physical well-being. 32 A study examining EAP impact on clinical outcomes demonstrated employees that accessed EAP services fared better clinically than ...

  16. Stress Factors in the Workplace Research Paper

    Conclusion. The paper has addressed the stress factors in the workplace and programs organizations could implement to help employee stress. For instance, it has been noted that toxic culture, bullying and harassment, relationships at work, and the death of patients and family are among the major stress factors for nurses.

  17. Workplace Stress and Productivity: A Cross-Sectional Study

    A multi-site, cross-sectional study was conducted to survey employees across four worksites participating in a WorkWell KS Well Being workshop to assess levels of stress and productivity. Stress was measured by the Perceived Stress Scale (PSS) and productivity was measured by the Health and Work Questionnaire (HWQ).

  18. Political Typology Quiz

    Take our quiz to find out which one of our nine political typology groups is your best match, compared with a nationally representative survey of more than 10,000 U.S. adults by Pew Research Center. You may find some of these questions are difficult to answer. That's OK. In those cases, pick the answer that comes closest to your view, even if ...

  19. Occupational Stress and Employees Complete Mental Health: A Cross

    Research on occupational stress can offer an essential function in the development of strategies aimed at the prevention and promotion of mental health at work . Thus, this study was the first to examine the relationship between occupational stress and complete mental health and the psychological mechanisms that account for such a relationship ...