[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (142)

Search Parameters:
Keywords = parabens

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 982 KiB  
Article
Comparative Perceptions of Fluoride Toxicity in Oral Hygiene Products: Insights from the General Population and Healthcare Professionals
by Marija Badrov, Lidia Gavic, Ana Seselja Perisin, Davor Zeljezic, Jasen Vladislavic, Ema Puizina Mladinic and Antonija Tadin
Clin. Pract. 2024, 14(5), 1827-1841; https://doi.org/10.3390/clinpract14050146 - 5 Sep 2024
Viewed by 363
Abstract
Background: The safety of oral hygiene products is a growing concern, particularly regarding the toxicity of specific ingredients used in their formulations. The purpose of this study was to evaluate the knowledge and attitudes of dentists, physicians, pharmacists, and the general public regarding [...] Read more.
Background: The safety of oral hygiene products is a growing concern, particularly regarding the toxicity of specific ingredients used in their formulations. The purpose of this study was to evaluate the knowledge and attitudes of dentists, physicians, pharmacists, and the general public regarding ingredients in oral hygiene products, especially fluoride. Additionally, this study aimed to identify which ingredients may exhibit potential toxicity based on historical records of any adverse effects being induced by a material/component. Methods: A self-administered questionnaire was used in an online cross-sectional observational study to collect data on sociodemographic characteristics, knowledge of fluoride in dental medicine, fluoride usage practices in oral hygiene products, opinions on ingredient toxicity in oral hygiene products, and personal experiences of adverse reactions to products and their components. The collected data underwent descriptive and regression analyses to reveal patterns and relationships within the dataset. Results: The study found a moderate overall knowledge level regarding fluoride usage in dentistry among participants (Md = 5.00, IQR 2.50–7.00). Healthcare professionals exhibited significantly higher knowledge scores compared to the general population (p ≤ 0.001), with dental professionals displaying the highest scores. Regarding concerns about the usage of fluoride, the majority of respondents (77.0%) did not express any concerns. Minor concerns included the risk of ingestion (6.0%) and dental fluorosis (4.6%). Among the other ingredients in oral hygiene products, respondents named alcohol as the most toxic ingredient (70.3%), followed by artificial colors (53.1%), artificial sweeteners (50.4%), and parabens (50.1%). It is noteworthy that the majority of participants (61.6%) stated that they had never experienced any side effects associated with the use of oral hygiene products. Conclusion: This study underscores disparities in fluoride knowledge between healthcare professionals and the general population in Croatia, with dental experts exhibiting a superior understanding. Despite lingering misconceptions about fluoride content and potential toxicity, the majority of participants acknowledge its oral health benefits and use fluoride products regularly. Full article
Show Figures

Figure 1

Figure 1
<p>Participant experiences of side effects associated with oral hygiene product usage.</p>
Full article ">Figure 2
<p>Oral hygiene products associated with potential oral mucosal damage as reported by participants.</p>
Full article ">Figure 3
<p>The distribution of primary concerns regarding fluoride usage among participants.</p>
Full article ">
16 pages, 3580 KiB  
Review
Synthetic Endocrine Disruptors in Fragranced Products
by Sawyer Ashcroft, Noura S. Dosoky, William N. Setzer and Prabodh Satyal
Endocrines 2024, 5(3), 366-381; https://doi.org/10.3390/endocrines5030027 - 15 Aug 2024
Viewed by 539
Abstract
Endocrine disruptors are molecules that can interfere with the proper functioning of the endocrine system and lead to harmful effects in living organisms. This review focuses on the impact of synthetic fragrances, which are commonly found in personal care and household products, on [...] Read more.
Endocrine disruptors are molecules that can interfere with the proper functioning of the endocrine system and lead to harmful effects in living organisms. This review focuses on the impact of synthetic fragrances, which are commonly found in personal care and household products, on the endocrine system. The article discusses the different types of hormones in the body and how they interact with receptors to produce signals. It also explores how endocrine disruptors can interfere with hormone signaling and transport, leading to adverse effects in the body. This work underscores the crucial need for further research into the impact of synthetic fragrances on the endocrine system and the importance of using safer alternatives in personal care and household products. Full article
(This article belongs to the Special Issue Feature Papers in Endocrines: 2024)
Show Figures

Figure 1

Figure 1
<p>A diagram of an external receptor. The hormone binds to the matching receptor on the outside of the cell, which triggers the release of a secondary messenger that produces a cellular response.</p>
Full article ">Figure 2
<p>A diagram of a nuclear receptor. The hormone binds to the matching receptor inside the cell, forming a dimer complex. The dimer then binds to DNA and transcribes an mRNA sequence. The mRNA then leaves the nucleus and is converted to a protein, which affects the function of the cell.</p>
Full article ">Figure 3
<p>Chemical structure of selected hormones. (<b>A</b>) Structure of prostaglandin E1, a type of eicosanoid used as a medication for infants with certain forms of heart defects to keep certain blood vessels open during surgery, (<b>B</b>) structure of estradiol, the major female sex hormone responsible for regulating healthy female development, (<b>C</b>) oxytocin is a peptide derivative responsible for childbirth and certain positive feedback loops in the brain, and (<b>D</b>) melatonin is an amino acid derivative used to regulate sleep cycles.</p>
Full article ">Figure 3 Cont.
<p>Chemical structure of selected hormones. (<b>A</b>) Structure of prostaglandin E1, a type of eicosanoid used as a medication for infants with certain forms of heart defects to keep certain blood vessels open during surgery, (<b>B</b>) structure of estradiol, the major female sex hormone responsible for regulating healthy female development, (<b>C</b>) oxytocin is a peptide derivative responsible for childbirth and certain positive feedback loops in the brain, and (<b>D</b>) melatonin is an amino acid derivative used to regulate sleep cycles.</p>
Full article ">Figure 4
<p>An example of how an endocrine-disrupting compound can mimic the structure of a hormone. The molecule on the left is estradiol, a natural hormone found in the body. The molecule on the right is an isomer of nonylphenol, a suspected EDC. The structural similarity is how a molecule other than a hormone can bind to a hormone receptor site despite the binding site’s specific structure.</p>
Full article ">Figure 5
<p>Chemical structures of endocrine-disrupting compounds in fragranced products. (<b>A</b>) The general structure of a phthalate. R represents a hydrocarbon moiety of any size. (<b>B</b>) The general structure of a paraben. R represents a hydrocarbon moiety of any size. (<b>C</b>) A diagram showing how one isomer of nonylphenol ethoxylate (left) can convert to nonylphenol (top right) or a shorter ethoxylate (bottom right). (<b>D</b>) Musk xylene is an example of a nitro musk. (<b>E</b>) HHCB, also known as Galaxolide, is an example of a polycyclic musk. (<b>F</b>) 2-hydroxy-4-methoxybenzophenone (oxybenzone), an organic blocker used as a UV filter. (<b>G</b>) Titanium dioxide, an inorganic blocker used as a UV filter. (<b>H</b>) Triclosan is a polychlorinated organic molecule. (<b>I</b>) Octamethylcyclotetrasiloxane (D4) is a common example of a cyclic methylsiloxane.</p>
Full article ">Figure 5 Cont.
<p>Chemical structures of endocrine-disrupting compounds in fragranced products. (<b>A</b>) The general structure of a phthalate. R represents a hydrocarbon moiety of any size. (<b>B</b>) The general structure of a paraben. R represents a hydrocarbon moiety of any size. (<b>C</b>) A diagram showing how one isomer of nonylphenol ethoxylate (left) can convert to nonylphenol (top right) or a shorter ethoxylate (bottom right). (<b>D</b>) Musk xylene is an example of a nitro musk. (<b>E</b>) HHCB, also known as Galaxolide, is an example of a polycyclic musk. (<b>F</b>) 2-hydroxy-4-methoxybenzophenone (oxybenzone), an organic blocker used as a UV filter. (<b>G</b>) Titanium dioxide, an inorganic blocker used as a UV filter. (<b>H</b>) Triclosan is a polychlorinated organic molecule. (<b>I</b>) Octamethylcyclotetrasiloxane (D4) is a common example of a cyclic methylsiloxane.</p>
Full article ">Figure 6
<p>A diagram that demonstrates bioaccumulation. EDCs can enter the body through several pathways and accumulate if the body cannot easily remove them.</p>
Full article ">Figure 7
<p>An example of biomagnification in an aquatic ecosystem. The green dots represent normal molecules present in the organism’s body, and the red symbols represent an EDC. As larger predators consume animals or plants lower on the food chain, they accumulate higher levels of EDCs from each level below.</p>
Full article ">Figure 8
<p>Linalool comes in two forms: (+)-linalool (<b>left</b>) or (-)-linalool (<b>right</b>). Each form has a different smell and a different biological activity.</p>
Full article ">
14 pages, 342 KiB  
Review
Use of Cosmetics in Pregnancy and Neurotoxicity: Can It Increase the Risk of Congenital Enteric Neuropathies?
by Kendra Jones, Lucas M. Wessel, Karl-Herbert Schäfer and María Ángeles Tapia-Laliena
Biomolecules 2024, 14(8), 984; https://doi.org/10.3390/biom14080984 - 10 Aug 2024
Viewed by 596
Abstract
Pregnancy is a particularly vulnerable period for the growing fetus, when exposure to toxic agents, especially in the early phases, can decisively harm embryo development and compromise the future health of the newborn. The inclusion of various chemical substances in personal care products [...] Read more.
Pregnancy is a particularly vulnerable period for the growing fetus, when exposure to toxic agents, especially in the early phases, can decisively harm embryo development and compromise the future health of the newborn. The inclusion of various chemical substances in personal care products (PCPs) and cosmetic formulations can be associated with disruption and damage to the nervous system. Microplastics, benzophenones, parabens, phthalates and metals are among the most common chemical substances found in cosmetics that have been shown to induce neurotoxic mechanisms. Although cosmetic neurotoxin exposure is believed to be minimal, different exposure scenarios of cosmetics suggest that these neurotoxins remain a threat. Special attention should be paid to early exposure in the first weeks of gestation, when critical processes, like the migration and proliferation of the neural crest derived cells, start to form the ENS. Importantly, cosmetic neurotoxins can cross the placental barrier and affect the future embryo, but they are also secreted in breast milk, so babies remain exposed for longer periods, even after birth. In this review, we explore how neurotoxins contained in cosmetics and PCPs may have a role in the pathogenesis of various neurodevelopmental disorders and neurodegenerative diseases and, therefore, also in congenital enteric aganglionosis as well as in postnatal motility disorders. Understanding the mechanisms of these chemicals used in cosmetic formulations and their role in neurotoxicity is crucial to determining the safety of use for cosmetic products during pregnancy. Full article
(This article belongs to the Special Issue Pathogenesis and Potential Treatments of Neurointestinal Diseases)
14 pages, 2589 KiB  
Article
Mixtures of Urinary Phenol and Phthalate Metabolite Concentrations in Relation to Serum Lipid Levels among Pregnant Women: Results from the EARTH Study
by Xilin Shen, Maximilien Génard-Walton, Paige L. Williams, Tamarra James-Todd, Jennifer B. Ford, Kathryn M. Rexrode, Antonia M. Calafat, Dan Zhang, Jorge E. Chavarro, Russ Hauser, Lidia Mínguez-Alarcón and the EARTH Study Team
Toxics 2024, 12(8), 574; https://doi.org/10.3390/toxics12080574 - 7 Aug 2024
Viewed by 595
Abstract
We examined whether mixtures of urinary concentrations of bisphenol A (BPA), parabens and phthalate metabolites were associated with serum lipid levels among 175 pregnant women who enrolled in the Environment and Reproductive Health (EARTH) Study (2005–2017), including triglycerides, total cholesterol, high-density lipoprotein (HDL), [...] Read more.
We examined whether mixtures of urinary concentrations of bisphenol A (BPA), parabens and phthalate metabolites were associated with serum lipid levels among 175 pregnant women who enrolled in the Environment and Reproductive Health (EARTH) Study (2005–2017), including triglycerides, total cholesterol, high-density lipoprotein (HDL), non-HDL, and low-density lipoprotein (LDL). We applied Bayesian Kernel Machine Regression (BKMR) and quantile g-computation while adjusting for confounders. In the BKMR models, we found no associations between chemical mixture and lipid levels, e.g., total cholesterol [mean difference (95% CRI, credible interval) = 0.02 (−0.31, 0.34)] and LDL [mean difference (95% CRI) = 0.10 (−0.22, 0.43)], when comparing concentrations at the 75th to the 25th percentile. When stratified by BMI, we found suggestive positive relationships between urinary propylparaben and total cholesterol and LDL among women with high BMI [mean difference (95% CRI) = 0.25 (−0.26, 0.75) and 0.35 (−0.25, 0.95)], but not with low BMI [mean difference (95% CRI) = 0.00 (−0.06, 0.07) and 0.00 (−0.07, 0.07)]. No association was found by quantile g-computation. This exploratory study suggests mixtures of phenol and phthalate metabolites were not associated with serum lipid levels during pregnancy, while there were some suggestive associations for certain BMI subgroups. Larger longitudinal studies with multiple assessments of both exposure and outcome are needed to corroborate these novel findings. Full article
(This article belongs to the Section Reproductive and Developmental Toxicity)
Show Figures

Figure 1

Figure 1
<p>Bayesian Kernel Machine Regression (BKMR) mixture associations of phenol and phthalate metabolites and lipid biomarkers. (<b>A</b>–<b>E</b>) Mixture associations with total triglycerides, total cholesterol, HDL, non-HDL and LDL cholesterol in BKMR. Upper half: exposure–response relationships for each biomarker while holding all other biomarkers at their medians. Lower half: mean difference between the 75th and 25th percentiles of exposure (estimates and 95% credible intervals) when other biomarker concentrations were fixed at the 25th, 50th, and 75th percentiles. BKMR models were adjusted for age, education level, race, infertility diagnosis, mode of conception, number of fetuses, trimester and specific gravity. HDL: high-density lipoprotein. LDL: low-density lipoprotein. MEHP: mono(2-ethylhexyl) phthalate. MEHHP: mono(2-ethyl-5-hydroxyhexyl) phthalate. MEOHP: mono(2-ethyl-5-oxohexyl) phthalate. MECPP: mono(2-ethyl-5-carboxypentyl). MBP: mono-n-butyl phthalate. MiBP: mono-isobutyl phthalate. MEP: monoethyl phthalate. MBzP: monobenzyl phthalate.</p>
Full article ">Figure 2
<p>Bayesian Kernel Machine Regression (BKMR) mixture associations of phenol and phthalate metabolites and lipid biomarkers when stratified by BMI. (<b>A</b>–<b>E</b>) Mixture associations with total triglycerides, total cholesterol, HDL, non-HDL and LDL cholesterol in BKMR. Upper half: exposure–response relationships for each biomarker while holding all other biomarkers at their median. Lower half: mean difference between the 75th and 25th percentile of exposure (estimates and 95% credible intervals) when other biomarker concentrations were fixed at the 25th, 50th, and 75th percentiles. BKMR models were adjusted for age, education level, race, infertility diagnosis, mode of conception, number of fetuses, trimester and specific gravity. HDL: high-density lipoprotein. LDL: low-density lipoprotein. MEHP: mono(2-ethylhexyl) phthalate. MEHHP: mono(2-ethyl-5-hydroxyhexyl) phthalate. MEOHP: mono(2-ethyl-5-oxohexyl) phthalate. MECPP: mono(2-ethyl-5-carboxypentyl). MBP: mono-n-butyl phthalate. MiBP: mono-isobutyl phthalate. MEP: monoethyl phthalate. MBzP: monobenzyl phthalate.</p>
Full article ">Figure 2 Cont.
<p>Bayesian Kernel Machine Regression (BKMR) mixture associations of phenol and phthalate metabolites and lipid biomarkers when stratified by BMI. (<b>A</b>–<b>E</b>) Mixture associations with total triglycerides, total cholesterol, HDL, non-HDL and LDL cholesterol in BKMR. Upper half: exposure–response relationships for each biomarker while holding all other biomarkers at their median. Lower half: mean difference between the 75th and 25th percentile of exposure (estimates and 95% credible intervals) when other biomarker concentrations were fixed at the 25th, 50th, and 75th percentiles. BKMR models were adjusted for age, education level, race, infertility diagnosis, mode of conception, number of fetuses, trimester and specific gravity. HDL: high-density lipoprotein. LDL: low-density lipoprotein. MEHP: mono(2-ethylhexyl) phthalate. MEHHP: mono(2-ethyl-5-hydroxyhexyl) phthalate. MEOHP: mono(2-ethyl-5-oxohexyl) phthalate. MECPP: mono(2-ethyl-5-carboxypentyl). MBP: mono-n-butyl phthalate. MiBP: mono-isobutyl phthalate. MEP: monoethyl phthalate. MBzP: monobenzyl phthalate.</p>
Full article ">
14 pages, 913 KiB  
Article
Association between United States Environmental Contaminants and the Prevalence of Psoriasis Derived from the National Health and Nutrition Examination Survey
by Linfen Guo, Beilin Tu, Deng Li, Lin Zhi, Yange Zhang, Haitao Xiao, Wei Li and Xuewen Xu
Toxics 2024, 12(7), 522; https://doi.org/10.3390/toxics12070522 - 19 Jul 2024
Viewed by 782
Abstract
(1) Background: Prolonged coexposure to environmental contaminants is reportedly associated with adverse impacts on skin health. However, the collective effects of contaminant mixtures on psoriasis prevalence remain unclear. (2) Methods: A nationally representative cohort study was conducted using data from the National Health [...] Read more.
(1) Background: Prolonged coexposure to environmental contaminants is reportedly associated with adverse impacts on skin health. However, the collective effects of contaminant mixtures on psoriasis prevalence remain unclear. (2) Methods: A nationally representative cohort study was conducted using data from the National Health and Nutrition Examination Survey 2003–2006 and 2009–2014. The association between contaminant exposures and psoriasis prevalence was analyzed through weighted quantile sum regressions, restricted cubic splines, and multivariable logistic regression. (3) Results: 16,453 participants and 60 contaminants in 8 groups were involved. After adjusting for demographics and comorbidities, exposure to urinary perchlorate, nitrate, and thiocyanate mixtures (OR: 1.10, 95% CI: 1.00–1.21) demonstrated a significant positive linear association with psoriasis prevalence. Ethyl paraben (OR: 1.21, 95% CI: 1.02–1.44) exhibited a significant positive correlation with psoriasis risk as an individual contaminant. The association between blood cadmium, lead, and mercury mixtures (OR: 1.10, 95% CI: 1.00–1.21), urinary perchlorate, nitrate, and thiocyanate mixtures (OR: 1.16, 95% CI: 1.00–1.34), and psoriasis prevalence was more pronounced in the lower healthy lifestyle score subgroup. (4) Conclusions: Exposure to perchlorate, nitrate, and thiocyanate mixtures, and ethyl paraben was associated with an elevated psoriasis prevalence. Furthermore, the association between cadmium and lead and mercury mixtures as well as perchlorate, nitrate and thiocyanate mixtures, and psoriasis prevalence was more pronounced in individuals with less healthy lifestyles. Full article
(This article belongs to the Section Human Toxicology and Epidemiology)
Show Figures

Figure 1

Figure 1
<p>Flow chart of participants who met inclusion criteria and were included in this study. Model I was adjusted for variables including age, sex, race, marital status, education, employment, family income-to-poverty ratio, health insurance, BMI, cigarette smoking, alcohol consumption, HEI-2015, and LTPA. Model II was further adjusted for self-reported arthritis, hypertension, diabetes, cancer, CVD, and stroke history based on Model I. BMI, body mass index; CVD, cardiovascular disease; HEI, healthy eating index; LTPA, leisure time physical activity.</p>
Full article ">Figure 2
<p>Restricted spline curves for correlations between environmental contaminant mixture groups (WQS indexes) and psoriasis prevalence (Model II cohort). (<b>a</b>) Arsenic species in urine; (<b>b</b>) cadmium, lead, and mercury in blood; (<b>c</b>) perchlorate, nitrate, and thiocyanate in urine; (<b>d</b>) phenols in urine; (<b>e</b>) phthalates in urine; (<b>f</b>) polyaromatic hydrocarbons in urine; (<b>g</b>) polyfluoroalkyl chemicals in blood; (<b>h</b>) pesticides in urine. The shaded areas depict data distribution, while the dotted lines correspond to the 95% confidence intervals. Model II was adjusted for age, sex, race, marital status, education, employment, family income-to-poverty ratio, health insurance, BMI, cigarette smoking, alcohol consumption, HEI-2015, LTPA, and self-reported history of arthritis, hypertension, diabetes, cancer, cardiovascular disease, and stroke.</p>
Full article ">
16 pages, 1750 KiB  
Article
Endocrine-Disrupting Chemicals Exposure and Neurocognitive Function in the General Population: A Community-Based Study
by Feng-Chieh Su, Yi-Chia Wei, Chiao-Yin Sun, Heng-Jung Hsu, Chin-Chan Lee, Yih-Ting Chen, Heng-Chih Pan, Cheng-Kai Hsu, Yun-An Liu and Chun-Yu Chen
Toxics 2024, 12(7), 514; https://doi.org/10.3390/toxics12070514 - 17 Jul 2024
Viewed by 816
Abstract
Background: Endocrine-disrupting chemicals (EDCs) are pervasive in everyday environments. The impacts of these chemicals, along with EDC-related lifestyle and dietary habits on neurocognitive function, are not well understood. Methods: The Chang Gung Community Medicine Research Center conducted a cross-sectional study involving 887 participants. [...] Read more.
Background: Endocrine-disrupting chemicals (EDCs) are pervasive in everyday environments. The impacts of these chemicals, along with EDC-related lifestyle and dietary habits on neurocognitive function, are not well understood. Methods: The Chang Gung Community Medicine Research Center conducted a cross-sectional study involving 887 participants. From this initial cohort, 120 individuals were selected based on their EDC exposure scores for detailed analysis. Among these, 67 participants aged 55 years or older were further chosen to undergo cognitive impairment assessments using the Ascertain Dementia-8 (AD-8) questionnaire. Results: These 67 older participants did not significantly differ in age, albuminuria, or estimated glomerular filtration rate compared to those with lower impairment scores. This study revealed that mono-(2-ethylhexyl) phthalate (MEHP) levels (8.511 vs. 6.432 µg/g creatinine, p = 0.038) were associated with greater risk of cognitive impairment (AD-8 ≥ 2). Statistical models adjusting for age, gender, and diabetes indicated that MEHP levels positively correlated with AD-8 scores, achieving statistical significance in more comprehensive models (β ± SE: 0.160 ± 0.076, p = 0.042). Logistic regression analysis underscored a significant positive association between high MEHP levels and higher AD-8 scores (odds ratio: 1.217, p = 0.006). Receiver operating characteristic curves highlighted the association of high MEHP levels and EDC exposure scores for significant cognitive impairment, with areas under the curve of 66.3% and 66.6%, respectively. Conclusion: Exposure to EDCs, specifically di-(2-ethylhexyl) phthalate, the precursor to MEHP, may be associated with neurocognitive impairment in middle-aged and older adults. Full article
Show Figures

Figure 1

Figure 1
<p>Flowchart of participants selected for urinary endocrine-disrupting chemicals estimation.</p>
Full article ">Figure 2
<p>Associations between urinary endocrine-disruption chemicals levels and Ascertain Dementia-8 scores. The blue dots represent individuals’ AD-8 scores versus the corresponding endocrine-disrupting chemical concentrations. The red line represents the regression line of the blue dots, and the red area represents the 95% confidence interval of the regression line. Abbreviations: BPA, bisphenol A; NP, nonylphenol; 4-t-OP, 4-tert-octylphenol; 2,4-di-t-BP, 2,4-di-tert-butylpheno; MP, methylparaben; EP, ethylparaben; PP, propylparaben; BP, butylparaben; MMP, monomethyl phthalate; MEP, monoethyl phthalate; MnBP, mono-(n-butyl) phthalate; MBzP, monobenzyl phthalate; MEHP, mono-(2-ethylhexyl) phthalate; MiNP, monoisononyl phthalate; BP-3, benzophenone-3.</p>
Full article ">Figure 3
<p>Receiver operating characteristic curves illustrating the performance of endocrine-disrupting chemicals (EDCs) to Ascertain Dementia-8 (AD-8) are presented. These include mono-(2-ethylhexyl) phthalate (MEHP) in (<b>a</b>), monobenzyl phthalate (MBzP) in (<b>b</b>), reversed hemoglobin in (<b>c</b>), EDC exposure score in (<b>d</b>), and the combined effect of MEHP and EDC exposure score in (<b>e</b>), obtained by multiplying both variables. These analyses focus on the development of Ascertain AD-8 scores of 2 or above, which indicate potential neurocognitive impairment. The blue dash lines represent diagonal reference lines.</p>
Full article ">
20 pages, 2871 KiB  
Article
Gender- and Obesity-Specific Association of Co-Exposure to Personal Care Product and Plasticizing Chemicals and Short Sleep Duration among Adults: Evidence from the National Health and Nutrition Examination Survey 2011–2016
by Francis Manyori Bigambo, Jian Sun, Chun Zhu, Songshan Zheng, Yang Xu, Di Wu, Yankai Xia and Xu Wang
Toxics 2024, 12(7), 503; https://doi.org/10.3390/toxics12070503 - 11 Jul 2024
Viewed by 783
Abstract
There is limited evidence about the gender- and obesity-specific effects of personal care product and plasticizing chemicals (PCPPCs) on short sleep duration in adults. We evaluated the gender- and obesity-specific association of co-exposure to PCPPCs and short sleep duration among adults aged 20–60 [...] Read more.
There is limited evidence about the gender- and obesity-specific effects of personal care product and plasticizing chemicals (PCPPCs) on short sleep duration in adults. We evaluated the gender- and obesity-specific association of co-exposure to PCPPCs and short sleep duration among adults aged 20–60 years using the National Health and Nutrition Examination Survey (NHANES) 2011–2016, a secondary data source from the United States. Seventeen PCPPCs, including five phenols, two parabens, and ten phthalates, were detected, and sleep duration was assessed among 3012 adults. Logistic regression, weighted quantile sum (WQS) regression, and Bayesian kernel machine regression (BKMR) were employed. We found that bisphenol A (BPA), mono (caboxy-isooctyl) phthalate (MCOP), and mono (3-carboxypropyl) phthalate (MCPP) were consistently positively associated with short sleep duration in both females and males regardless of obesity status, except for BPA with general obesity. In particular, mono benzyl phthalate (MBzP) revealed a positive association in females, mono (2-ethyl-5-oxohexyl) phthalate (MEOHP) revealed a positive association in males, and MiBP revealed a positive association in abdominal obesity. Similar associations were observed in the mixture. Our study highlights that PCPPCs are independently associated with an increasing risk of short sleep duration in adults both individually and as a mixture; however, gender- and obesity-specific differences may have little effect on certain individual PCPPCs on short sleep duration. Full article
(This article belongs to the Section Exposome Analysis and Risk Assessment)
Show Figures

Figure 1

Figure 1
<p>Weighted quantile sum (WQS) regression model index weight for a mixture of personal care product and plasticizing chemicals (PCPPCs) with short sleep duration among adults by gender-specific status. (<b>A</b>) Female. (<b>B</b>) Male. The red line indicates the cutoff point to discriminate the relative importance of the chemical. The model was adjusted for age, race, education, marital status, body mass index (BMI) [not adjusted for general obesity and no general obesity], waist circumference (not adjusted for abdominal obesity and no abdominal obesity), family income to poverty ratio (PIR), food insecurity, physical activity, log<sub>10</sub> cotinine, and log<sub>10</sub> creatinine.</p>
Full article ">Figure 2
<p>Weighted quantile sum (WQS) regression model index weight for a mixture of personal care product and plasticizing chemicals (PCPPCs) with short sleep duration among adults by obesity-specific status. (<b>A</b>) General obesity. (<b>B</b>) No general obesity. (<b>C</b>) Abdominal obesity. (<b>D</b>) No abdominal obesity. The red line indicates the cutoff point to discriminate the relative importance of the chemical. The model was adjusted for age, gender, race, education, marital status, body mass index (BMI) [not adjusted for general obesity and no general obesity], waist circumference (not adjusted for abdominal obesity and no abdominal obesity), family income to poverty ratio (PIR), food insecurity, physical activity, log<sub>10</sub> cotinine, and log<sub>10</sub> creatinine.</p>
Full article ">Figure 3
<p>Univariate exposure–outcome function (95% credible interval) for personal care product and plasticizing chemicals (PCPPCs) with short sleep duration among adults by gender- and obesity-specific status. The function was estimated while fixing other PCPPCs at their median level. (<b>A</b>) Female. (<b>B</b>) Male. (<b>C</b>) General obesity. (<b>D</b>) No general obesity. (<b>E</b>) Abdominal obesity. (<b>F</b>) No abdominal obesity. The Bayesian kernel machine regression model was adjusted for age, gender (not adjusted for female and male), race, education, marital status, body mass index (BMI) [not adjusted for general obesity and no general obesity], waist circumference (not adjusted for abdominal obesity and no abdominal obesity), family income to poverty ratio (PIR), food insecurity, physical activity, log<sub>10</sub> cotinine, and log<sub>10</sub> creatinine.</p>
Full article ">Figure 4
<p>The mixture effect (95% confidence interval) of co-exposure to 17 personal care product and plasticizing chemicals (PCPPCs) on short sleep duration among adults by gender- and obesity-specific status. (<b>A</b>) Female. (<b>B</b>) Male. (<b>C</b>) General obesity. (<b>D</b>) No general obesity. (<b>E</b>) Abdominal obesity. (<b>F</b>) No abdominal obesity. The Bayesian kernel machine regression model was adjusted for age, gender (not adjusted for female and male), race, education, marital status, body mass index (BMI) [not adjusted for general obesity and no general obesity], waist circumference (not adjusted for abdominal obesity and no abdominal obesity), family income to poverty ratio (PIR), food insecurity, physical activity, log<sub>10</sub> cotinine, and log<sub>10</sub> creatinine.</p>
Full article ">
22 pages, 1978 KiB  
Article
A Personalized Intervention to Increase Environmental Health Literacy and Readiness to Change in a Northern Nevada Population: Effects of Environmental Chemical Exposure Report-Back
by Johanna R. Rochester, Carol F. Kwiatkowski, Iva Neveux, Shaun Dabe, Katherine M. Hatcher, Michael Kupec Lathrop, Eric J. Daza, Brenda Eskenazi, Joseph J. Grzymski and Jenna Hua
Int. J. Environ. Res. Public Health 2024, 21(7), 905; https://doi.org/10.3390/ijerph21070905 - 11 Jul 2024
Viewed by 1029
Abstract
Background: Interventions are needed to help people reduce exposure to harmful chemicals from everyday products and lifestyle habits. Report-back of individual exposures is a potential pathway to increasing environmental health literacy (EHL) and readiness to reduce exposures. Objectives: Our objective was to determine [...] Read more.
Background: Interventions are needed to help people reduce exposure to harmful chemicals from everyday products and lifestyle habits. Report-back of individual exposures is a potential pathway to increasing environmental health literacy (EHL) and readiness to reduce exposures. Objectives: Our objective was to determine if report-back of endocrine-disrupting chemicals (EDCs) can reduce EDC exposure, increase EHL, and increase readiness to change (i.e., to implement EDC exposure-reduction behaviors). Methods: Participants in the Healthy Nevada Project completed EHL and readiness-to-change surveys before (n = 424) and after (n = 174) a report-back intervention. Participants used mail-in kits to measure urinary biomarkers of EDCs. The report-back of results included urinary levels, information about health effects, sources of exposure, and personalized recommendations to reduce exposure. Results: EHL was generally very high at baseline, especially for questions related to the general pollution. For questions related to chemical exposures, responses varied across several demographics. Statistically reliable improvements in EHL responses were seen after report-back. For readiness to change, 72% were already or planning to change their behaviors. Post-intervention, women increased their readiness (p = 0.053), while men decreased (p = 0.007). When asked what challenges they faced in reducing exposure, 79% cited not knowing what to do. This dropped to 35% after report-back. Participants with higher propylparaben were younger (p = 0.03) and women and participants who rated themselves in better health had higher levels of some phthalates (p = 0.02–0.003 and p = 0.001–0.003, respectively). After report-back, monobutyl phthalate decreased among the 48 participants who had valid urine tests before and after the intervention (p < 0.001). Conclusions: The report-back intervention was successful as evidenced by increased EHL behaviors, increased readiness to change among women, and a decrease in monobutyl phthalate. An EHL questionnaire more sensitive to chemical exposures would help differentiate high and low literacy. Future research will focus on understanding why men decreased their readiness to change and how the intervention can be improved for all participants. Full article
(This article belongs to the Special Issue Research on Environmental Exposure, Pollution, and Epidemiology)
Show Figures

Figure 1

Figure 1
<p>Participation flow chart.</p>
Full article ">Figure 2
<p>Readiness to change by sex (<span class="html-italic">n</span> = 432).</p>
Full article ">Figure 3
<p>Number of participants responding to pre- and post-test “Readiness to Change” questions (<span class="html-italic">n</span> = 172).</p>
Full article ">Figure 4
<p>Pre-post changes in “Readiness to Change” for participants who completed post-test surveys (<span class="html-italic">n</span> = 172).</p>
Full article ">Figure 5
<p>Challenges to reducing exposure to harmful chemicals.</p>
Full article ">
21 pages, 4388 KiB  
Article
Microspheres Based on Blends of Chitosan Derivatives with Carrageenan as Vitamin Carriers in Cosmeceuticals
by Kamila Lewicka, Anna Smola-Dmochowska, Piotr Dobrzyński, Natalia Śmigiel-Gac, Katarzyna Jelonek, Monika Musiał-Kulik and Piotr Rychter
Polymers 2024, 16(13), 1815; https://doi.org/10.3390/polym16131815 - 26 Jun 2024
Viewed by 1254
Abstract
Chitosan (CS) has a natural origin and is a biodegradable and biocompatible polymer with many skin-beneficial properties successfully used in the cosmetics and pharmaceutical industry. CS derivatives, especially those synthesized via a Schiff base reaction, are very important due to their unique antimicrobial [...] Read more.
Chitosan (CS) has a natural origin and is a biodegradable and biocompatible polymer with many skin-beneficial properties successfully used in the cosmetics and pharmaceutical industry. CS derivatives, especially those synthesized via a Schiff base reaction, are very important due to their unique antimicrobial activity. This study demonstrates research results on the use of hydrogel microspheres made of [chitosan-graft-poly(ε-caprolactone)]-blend-(ĸ-carrageenan)], [chitosan-2-pyridinecarboxaldehyde-graft-poly(ε-caprolactone)]-blend-(ĸ-carrageenan), and chitosan-sodium-4-formylbenzene-1,3-disulfonate-graft-poly(ε-caprolactone)]-blend-(ĸ-carrageenan) as innovative vitamin carriers for cosmetic formulation. A permeation study of retinol (vitamin A), L-ascorbic acid (vitamin C), and α-tocopherol (vitamin E) from the cream through a human skin model by the Franz Cell measurement system was presented. The quantitative analysis of the release of the vitamins added to the cream base, through the membrane, imitating human skin, showed a promising profile of its release/penetration, which is promising for the development of a cream with anti-aging properties. Additionally, the antibacterial activity of the polymers from which the microspheres are made allows for the elimination of preservatives and parabens as cosmetic formulation ingredients. Full article
Show Figures

Figure 1

Figure 1
<p>Scanning electron microscopy (SEM) images of microspheres CS-g-PCL(MSA):CG 50:50 loaded with VA vitamins (<b>a</b>) after receiving and (<b>b</b>) swelling microspheres after 12 h of incubation in apH 5.0 buffer.</p>
Full article ">Figure 2
<p>Scanning electron microscopy (SEM) images of microspheres CS-SB-PCA-g-PCL:CG 50:50 loaded with VA vitamins (<b>a</b>) after receiving and (<b>b</b>) swelling microspheres after 12 h of incubation in apH 5.0 buffer.</p>
Full article ">Figure 3
<p>Scanning electron microscopy (SEM) images of microspheres CS-SB-SFD-g-PCL:CG 50:50 loaded with VA vitamins (<b>a</b>) after receiving and (<b>b</b>) swelling microspheres after 12 h of incubation in a pH 5.0 buffer.</p>
Full article ">Figure 4
<p>The diameter distribution of the microspheres loaded with vitamins after receiving (before swelling) and after 12 h of incubation in a pH 5.0 buffer.</p>
Full article ">Figure 5
<p>Swelling ratio (SR) of microspheres at different soaking times in a pH 5.0 buffer at 34 °C. Error bars represent the standard deviation of three replicates.</p>
Full article ">Figure 6
<p>The relative mass loss (M<sub>s</sub>%) of microspheres after the incubation process in a pH 5.0 buffer at 34 °C. Error bars represent the standard deviation of three replicates.</p>
Full article ">Figure 7
<p>VA permeated through the Strat-MTM membrane: (<b>a</b>) 1 h doses and (<b>b</b>) cumulative release. Error bars represent the standard deviation of three replicates.</p>
Full article ">Figure 8
<p>VC permeated through the Strat-MTM membrane: (<b>a</b>) 1 h doses and (<b>b</b>) cumulative release. Error bars represent the the standard deviation of three replicates.</p>
Full article ">Figure 9
<p>VE permeated through the Strat-MTM membrane: (<b>a</b>) 1 h doses and (<b>b</b>) cumulative release. Error bars represent the standard deviation of three replicates.</p>
Full article ">Figure 10
<p>Relative vitamin content in the cream, membrane, and permeate after 6 h of the release process of VA (<b>a</b>), VC (<b>b</b>) and VE (<b>c</b>).</p>
Full article ">Figure 11
<p>The effect of the developed blends on the viability of human fibroblasts (the results are shown as the mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 compared with the control).</p>
Full article ">Figure 12
<p>The effect of the developed blends on the viability of human keratinocytes (the results are shown as the mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 compared with the control).</p>
Full article ">
15 pages, 3709 KiB  
Article
A Survey of Preservatives Used in Cosmetic Products
by Patrycja Poddębniak and Urszula Kalinowska-Lis
Appl. Sci. 2024, 14(4), 1581; https://doi.org/10.3390/app14041581 - 16 Feb 2024
Cited by 3 | Viewed by 4382
Abstract
The aim of this study was to indicate the type of preservatives used in selected categories of cosmetic products sold in Poland (part of the EU market) and determine the frequency of their use. The tested products consisted of 200 leave-on cosmetics, viz. [...] Read more.
The aim of this study was to indicate the type of preservatives used in selected categories of cosmetic products sold in Poland (part of the EU market) and determine the frequency of their use. The tested products consisted of 200 leave-on cosmetics, viz. body lotions (n = 100) and face creams (n = 100) and rinse-off cosmetics (n = 100) and mascaras (n = 25). The product labels of 325 adult cosmetic products from international brands were analyzed for the presence of preservatives based on the INCI compositions. The survey focused on preservatives included in Annex V of the Regulation (EC) No. 1223/2009 of the European Parliament and Council of 30 November 2009 on cosmetic products. The tested products contained 29 different preservatives belonging to eight chemical groups. Most preservatives were alcohols or their derivatives, carboxylic acids or their salts, or parabens. The most common types were phenoxyethanol, present in 198/325 (60.9%) formulations, followed by sodium benzoate, in 137 (42.2%), potassium sorbate, in 116 (35.7%), benzyl alcohol, in 76 (23.4%), and methylparaben in 33 (10.2%). Also, 33 of the 60 preservatives included in Annex V of Regulation (EC) No. 1223/2009 were not used in any of the tested preparations. In each category of products, the most common were combinations of two preservatives per single product (34.8% of all products), followed by single-preservative products (25.5%) and three-preservative products (19.4%). Full article
(This article belongs to the Special Issue Development of Innovative Cosmetics)
Show Figures

Figure 1

Figure 1
<p>Frequency of use of preservatives in body lotions (n = 100).</p>
Full article ">Figure 2
<p>Frequency of use of preservatives in face creams (n = 100).</p>
Full article ">Figure 3
<p>Frequency of use of preservatives in rinse-off cosmetics (n = 100).</p>
Full article ">Figure 4
<p>Frequency of use of preservatives in mascaras (n = 25).</p>
Full article ">Figure 5
<p>Number of occurrences of the preservatives identified in the tested products (n = 325).</p>
Full article ">Figure 6
<p>Frequency of use of preservatives in single-preservative products (n = 83).</p>
Full article ">Figure 7
<p>Frequency of use of preservatives in two-preservative products (n = 113).</p>
Full article ">Figure 8
<p>Prevalence of preservatives in body lotions (n = 100) from ten different cosmetic brands.</p>
Full article ">Figure 9
<p>Prevalence of preservatives in face creams (n = 100) from ten different cosmetic brands.</p>
Full article ">Figure 10
<p>Prevalence of preservatives in rinse-off products (n = 100) from ten different cosmetic brands.</p>
Full article ">
13 pages, 1244 KiB  
Article
Multiclass Determination of Endocrine-Disrupting Chemicals in Meconium: First Evidence of Perfluoroalkyl Substances in This Biological Compartment
by Aritz Domínguez-Liste, Teresa de Haro-Romero, Raquel Quesada-Jiménez, Ainhoa Pérez-Cantero, Francisco Manuel Peinado, Óscar Ballesteros and Fernando Vela-Soria
Toxics 2024, 12(1), 75; https://doi.org/10.3390/toxics12010075 - 15 Jan 2024
Cited by 1 | Viewed by 1403
Abstract
Major concerns have been raised about human exposure to endocrine-disrupting chemicals (EDCs) during pregnancy. Effective methodologies for the assessment of this exposure are needed to support the implementation of preventive measures and the prediction of negative health effects. Meconium has proven a valuable [...] Read more.
Major concerns have been raised about human exposure to endocrine-disrupting chemicals (EDCs) during pregnancy. Effective methodologies for the assessment of this exposure are needed to support the implementation of preventive measures and the prediction of negative health effects. Meconium has proven a valuable non-invasive matrix for evaluating cumulative exposure to xenobiotics during the last two trimesters of pregnancy. The study objective was to develop a novel method to determine the presence in meconium of perfluoroalkyl substances (PFASs), bisphenols, parabens, and benzophenones, EDCs that are widely used in the manufacture of numerous consumer goods and personal care products, including cosmetics. Ten PFASs, two bisphenols, four parabens, and four benzophenones were measured in meconium samples prepared by using a combination of Captiva Enhanced Matrix Removal (EMR) lipid cartridges with salt-assisted liquid–liquid extraction (SALLE) and dispersive liquid–liquid microextraction (DLLME) before the application of liquid chromatography–tandem mass spectrometry (LC–MS/MS). Experimental parameters were optimized by applying different chemometric techniques. Limits of detection ranged from 0.05 to 0.1 ng g−1, and between-day variabilities (relative standard deviations) ranged from 6.5% to 14.5%. The method was validated by matrix-matched standard calibration followed by a recovery assay with spiked samples, obtaining percentage recoveries of 89.9% to 114.8%. The method was then employed to measure compounds not previously studied in this matrix in 20 meconium samples. The proposed analytical procedure yields information on cumulative in utero exposure to selected EDCs. Full article
(This article belongs to the Section Emerging Contaminants)
Show Figures

Figure 1

Figure 1
<p>Estimated response surface for desirability function in SALLE optimization.</p>
Full article ">Figure 2
<p>Pareto diagrams showing the significance of standardized effects of DLLME conditions on peak areas for PFTrA, PFHxS, and BP-1.</p>
Full article ">
14 pages, 1843 KiB  
Article
Elucidation of 4-Hydroxybenzoic Acid Catabolic Pathways in Pseudarthrobacter phenanthrenivorans Sphe3
by Epameinondas Tsagogiannis, Stamatia Asimakoula, Alexandros P. Drainas, Orfeas Marinakos, Vasiliki I. Boti, Ioanna S. Kosma and Anna-Irini Koukkou
Int. J. Mol. Sci. 2024, 25(2), 843; https://doi.org/10.3390/ijms25020843 - 10 Jan 2024
Cited by 1 | Viewed by 1254
Abstract
4-hydroxybenzoic acid (4-HBA) is an aromatic compound with high chemical stability, being extensively used in food, pharmaceutical and cosmetic industries and therefore widely distributed in various environments. Bioremediation constitutes the most sustainable approach for the removal of 4-hydroxybenzoate and its derivatives (parabens) from [...] Read more.
4-hydroxybenzoic acid (4-HBA) is an aromatic compound with high chemical stability, being extensively used in food, pharmaceutical and cosmetic industries and therefore widely distributed in various environments. Bioremediation constitutes the most sustainable approach for the removal of 4-hydroxybenzoate and its derivatives (parabens) from polluted environments. Pseudarthrobacter phenanthrenivorans Sphe3, a strain capable of degrading several aromatic compounds, is able to grow on 4-HBA as the sole carbon and energy source. Here, an attempt is made to clarify the catabolic pathways that are involved in the biodegradation of 4-hydroxybenzoate by Sphe3, applying a metabolomic and transcriptomic analysis of cells grown on 4-HBA. It seems that in Sphe3, 4-hydroxybenzoate is hydroxylated to form protocatechuate, which subsequently is either cleaved in ortho- and/or meta-positions or decarboxylated to form catechol. Protocatechuate and catechol are funneled into the TCA cycle following either the β-ketoadipate or protocatechuate meta-cleavage branches. Our results also suggest the involvement of the oxidative decarboxylation of the protocatechuate peripheral pathway to form hydroxyquinol. As a conclusion, P. phenanthrenivorans Sphe3 seems to be a rather versatile strain considering the 4-hydroxybenzoate biodegradation, as it has the advantage to carry it out effectively following different catabolic pathways concurrently. Full article
(This article belongs to the Collection Feature Papers in Molecular Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>Graphical representation of growth profiles and 4-HBA biodegradation by <span class="html-italic">P. phenanthrenivorans</span> Sphe3 and Sphe3c strains. The growth of both strains was measured using optical density at 600 nm while biodegradation rate of 4-HBA in the culture medium by the respective strains was measured using HPLC, expressed as percentage (%) of remaining 4-HBA concentration at various time points.</p>
Full article ">Figure 2
<p>Map of pASPHE301 plasmid from <span class="html-italic">P. phenanthrenivorans</span> Sphe3. Marked with red circles are catabolic genes that concern the present study. 1: gentisate 1,2-dioxygenase (<span class="html-italic">gdo</span>12, Asphe3_39840), 2: catechol 2,3-dioxygenase (<span class="html-italic">cdo</span>23, Asphe3_40510). The plasmid map was extracted from the Joint Genomic Institute (JGI) database (<a href="https://img.jgi.doe.gov/cgi-bin/m/main.cgi?section=TaxonDetail&amp;page=taxonDetail&amp;taxon_oid=650377905#genomedetail2" target="_blank">https://img.jgi.doe.gov/cgi-bin/m/main.cgi?section=TaxonDetail&amp;page=taxonDetail&amp;taxon_oid=650377905#genomedetail2</a>) (accessed on 22 July 2023).</p>
Full article ">Figure 3
<p>Transcription quantification for the genes involved in the biodegradation of 4-HBA in <span class="html-italic">P. phenenthrenivorans</span> Sphe3. The studied genes are coding for catechol 1,2- and 2,3- dioxygenases (<span class="html-italic">cdo</span>12 and <span class="html-italic">cdo</span>23, respectively) and PCA 3,4- and 4,5-dioxygenases (<span class="html-italic">pcd</span>34 and <span class="html-italic">pcd</span>45) monitored using RT-qPCR in Sphe3 and Sphe3c cells grown on glucose and 4-HBA, each as the sole carbon and energy source. Values represent the mean relative gene expression normalized to the housekeeping gene <span class="html-italic">gyr</span>β ± standard deviations of three individual replicates. Gene expression levels in glucose were used as a calibrator.</p>
Full article ">Figure 4
<p>Proposed pathways for the catabolism of 4-HBA in <span class="html-italic">P. phenanthrenivorans</span> Sphe3 according to the metabolite analysis using LC-MS. Enzymes expected to catalyze the respective reactions (detected in the Sphe3 genome after the in silico analysis) are mentioned: 4HB3H, 4-hydroxybenzoate-3-hydroxylase; PCD45, PCA 4,5-dioxygenase; PCD34, PCA 3,4- dioxygenase; CDO12, catechol 1,2-dioxygenase; CDO23, catechol 2,3-dioxygenase. All the products are funneled into the central metabolism of the cell (TCA cycle).</p>
Full article ">
14 pages, 1950 KiB  
Article
Parabens, Triclosan and Bisphenol A in Surface Waters and Sediments of Baiyang Lake, China: Occurrence, Distribution, and Potential Risk Assessment
by Liguo Fu, Yaxue Sun, Jingbo Zhou, Hongbo Li and Shu-xuan Liang
Toxics 2024, 12(1), 31; https://doi.org/10.3390/toxics12010031 - 31 Dec 2023
Cited by 2 | Viewed by 1424
Abstract
The extensive use of the parabens triclosan (TCS) and bisphenol A (BPA) has potential adverse effects on human health and aquatic organisms. However, their monitoring information in freshwater lakes is still limited. This study simultaneously summarized the concentrations, spatial distribution characteristics, and correlations [...] Read more.
The extensive use of the parabens triclosan (TCS) and bisphenol A (BPA) has potential adverse effects on human health and aquatic organisms. However, their monitoring information in freshwater lakes is still limited. This study simultaneously summarized the concentrations, spatial distribution characteristics, and correlations of four types of parabens, TCS, and BPA in the surface water and sediment of Baiyang Lake. Finally, the potential risks of target pollutants were evaluated from two aspects: human health risks and ecological risks. The average contaminations of target compounds in surface water and sediment—BPA, TCS, and ∑4 parabens—was 33.1, 26.1, 0.7 ng/L and 24.5, 32.5, 2.5 ng/g, respectively. The total concentration of target compounds at the inlet of the upstream Fu River and Baigouyin River is significantly higher than that near Hunan and the outlet. In addition, Spearman’s correlation analysis showed a significant positive correlation between compounds. The health hazards of target compounds in surface water were all within safe limits. However, the risk quotient results indicate that in some locations in surface water, TCS poses a high risk to algae and a moderate risk to invertebrates and fish, and appropriate attention should be paid to these areas. Full article
Show Figures

Figure 1

Figure 1
<p>Map of sampling sites in Baiyang Lake.</p>
Full article ">Figure 2
<p>Spatial distribution of total parabens, TCS and BPA in surface water (<b>a</b>) and sediments (<b>b</b>).</p>
Full article ">Figure 3
<p>Spatial distribution of BPA (<b>a</b>,<b>d</b>), TCS (<b>b</b>,<b>e</b>), and parabens monomer (<b>c</b>,<b>f</b>) in surface water and sediments.</p>
Full article ">Figure 4
<p>Overall and mean risk quotient of five target analytes in surface water at different sites for three groups of peoples.</p>
Full article ">Figure 5
<p>Ecological risk assessment of the target compounds at different trophic levels of aquatic organisms: algae, crustaceans, and fish in surface water, (<b>a</b>) and algae, crustaceans, and fish in sediment (<b>b</b>).</p>
Full article ">
12 pages, 1942 KiB  
Article
Development and Characterization of Econazole Topical Gel
by Mohammad F. Bayan, Balakumar Chandrasekaran and Mohammad H. Alyami
Gels 2023, 9(12), 929; https://doi.org/10.3390/gels9120929 - 25 Nov 2023
Cited by 5 | Viewed by 2454
Abstract
The purpose of this work was to develop a novel topical formulation of econazole nitrate based on gel that can be easily scaled up in one pot for the potential treatment of fungal and yeast infections. Econazole nitrate, a topical antifungal, is used [...] Read more.
The purpose of this work was to develop a novel topical formulation of econazole nitrate based on gel that can be easily scaled up in one pot for the potential treatment of fungal and yeast infections. Econazole nitrate, a topical antifungal, is used to treat tinea versicolor, tinea pedis, and tinea cruris. Compared to applying cream or ointment, topical gels offer numerous advantages, one of which is that the drug is released more quickly to the intended site of action. A viscous mixture of propylene glycol, Capmul® MCM C8, methyl and propyl paraben, and econazole nitrate were mixed together before being formulated into the optimized Carbopol® gel bases. The gel’s color, appearance, and homogeneity were assessed visually. For every formulation, the drug content, pH, viscosity, spreadability, and gel strength were characterized. The cup plate diffusion method was used to evaluate the anti-fungal activity of the prepared formulations. To assess the behavior of the developed system, studies on in vitro release and mechanism were conducted. The manufactured formulations were transparent, pale yellow, and exhibited excellent homogeneity. The pH of each formulation was roughly 6.0, making them suitable for topical use. The concentration of Carbopol® 940 resulted in a significant increase in viscosity and gel strength but a significant decrease in spreadability. It was demonstrated that the prepared formulations inhibited the growth of Candida albicans and Aspergillus fumigatus. In contrast, the standard blank gel showed no signs of antifungal action. By increasing the concentration of Carbopol® 940, the in vitro release profile of econazole nitrate significantly decreased. Following the Korsmeyer–Peppas model fitting, all formulations exhibited n values greater than 0.5 and less than 1, indicating that diffusion and gel swelling control econazole nitrate release. Full article
(This article belongs to the Special Issue Advanced Gels for Wound Healing)
Show Figures

Figure 1

Figure 1
<p>The effect of Carbopol<sup>®</sup> 940 concentration on viscosity.</p>
Full article ">Figure 2
<p>The effect of Carbopol<sup>®</sup> 940 concentration on spreadability.</p>
Full article ">Figure 3
<p>The effect of Carbopol<sup>®</sup> 940 concentration on gel strength.</p>
Full article ">Figure 4
<p>Econzaole nitrate in vitro release profiles.</p>
Full article ">Figure 5
<p>Korsmeyer–Peppas model release kinetics of econazole nitrate formulations.</p>
Full article ">Figure 6
<p>Calibration curve of econzaole nitrate.</p>
Full article ">
17 pages, 374 KiB  
Article
Expression Profiles of Genes Related to Development and Progression of Endometriosis and Their Association with Paraben and Benzophenone Exposure
by Francisco M. Peinado, Alicia Olivas-Martínez, Inmaculada Lendínez, Luz M. Iribarne-Durán, Josefa León, Mariana F. Fernández, Rafael Sotelo, Fernando Vela-Soria, Nicolás Olea, Carmen Freire, Olga Ocón-Hernández and Francisco Artacho-Cordón
Int. J. Mol. Sci. 2023, 24(23), 16678; https://doi.org/10.3390/ijms242316678 - 23 Nov 2023
Cited by 1 | Viewed by 1134
Abstract
Increasing evidence has been published over recent years on the implication of endocrine-disrupting chemicals (EDCs), including parabens and benzophenones in the pathogenesis and pathophysiology of endometriosis. However, to the best of our knowledge, no study has been published on the ways in which [...] Read more.
Increasing evidence has been published over recent years on the implication of endocrine-disrupting chemicals (EDCs), including parabens and benzophenones in the pathogenesis and pathophysiology of endometriosis. However, to the best of our knowledge, no study has been published on the ways in which exposure to EDCs might affect cell-signaling pathways related to endometriosis. We aimed to describe the endometriotic tissue expression profile of a panel of 23 genes related to crucial cell-signaling pathways for the development and progression of endometriosis (cell adhesion, invasion/migration, inflammation, angiogenesis, and cell proliferation/hormone stimulation) and explore its relationship with the exposure of patients to parabens (PBs) and benzophenones (BPs). This cross-sectional study included a subsample of 33 women with endometriosis from the EndEA study, measuring their endometriotic tissue expressions of 23 genes, while urinary concentrations of methyl-, ethyl-, propyl-, butyl-paraben, benzophenone-1, benzophenone-3, and 4-hydroxybenzophenone were determined in 22 women. Spearman’s correlations test and linear and logistic regression analyses were performed. The expression of 52.2% of studied genes was observed in >75% of endometriotic tissue samples and the expression of 17.4% (n = 4) of them in 50–75%. Exposure to certain PB and BP congeners was positively associated with the expression of key genes for the development and proliferation of endometriosis. Genes related to the development and progression of endometriosis were expressed in most endometriotic tissue samples studied, suggesting that exposure of women to PBs and BPs may be associated with the altered expression profile of genes related to cellular pathways involved in the development of endometriosis. Full article
(This article belongs to the Section Molecular Biology)
Back to TopTop