Diet and Immune Function
Diet and Immune Function
Diet and Immune Function
Diet and
Immune Function
Edited by
Elizabeth A. Miles, Philip Calder and Caroline E. Childs
Printed Edition of the Special Issue Published in Nutrients
www.mdpi.com/journal/nutrients
Diet and Immune Function
Diet and Immune Function
Caroline E Childs
University of Southampton
UK
Editorial Office
MDPI
St. Alban-Anlage 66
4052 Basel, Switzerland
This is a reprint of articles from the Special Issue published online in the open access journal Nutrients
(ISSN 2072-6643) from 2018 to 2019 (available at: https://www.mdpi.com/journal/nutrients/
special issues/diet-immune)
For citation purposes, cite each article independently as indicated on the article page online and as
indicated below:
LastName, A.A.; LastName, B.B.; LastName, C.C. Article Title. Journal Name Year, Article Number,
Page Range.
c 2020 by the authors. Articles in this book are Open Access and distributed under the Creative
Commons Attribution (CC BY) license, which allows users to download, copy and build upon
published articles, as long as the author and publisher are properly credited, which ensures maximum
dissemination and a wider impact of our publications.
The book as a whole is distributed by MDPI under the terms and conditions of the Creative Commons
license CC BY-NC-ND.
Contents
Bethan Dalton, Iain C. Campbell, Raymond Chung, Gerome Breen, Ulrike Schmidt and
Hubertus Himmerich
Inflammatory Markers in Anorexia Nervosa: An Exploratory Study
Reprinted from: Nutrients 2018, 10, 1573, doi:10.3390/nu10111573 . . . . . . . . . . . . . . . . . . 10
Starin McKeen, Wayne Young, Jane Mullaney, Karl Fraser, Warren C. McNabb and
Nicole C. Roy
Infant Complementary Feeding of Prebiotics for the Microbiome and Immunity
Reprinted from: Nutrients 2019, 11, 364, doi:10.3390/nu11020364 . . . . . . . . . . . . . . . . . . . 55
Vinicius Cruzat, Marcelo Macedo Rogero, Kevin Noel Keane, Rui Curi and
Philip Newsholme
Glutamine: Metabolism and Immune Function, Supplementation and Clinical Translation
Reprinted from: Nutrients 2018, 10, 1564, doi:10.3390/nu10111564 . . . . . . . . . . . . . . . . . . 151
v
Silvia Maggini, Adeline Pierre and Philip C. Calder
Immune Function and Micronutrient Requirements Change over the Life Course
Reprinted from: Nutrients 2018, 10, 1531, doi:10.3390/nu10101531 . . . . . . . . . . . . . . . . . . 182
vi
About the Special Issue Editors
Elizabeth A Miles is Lecturer in Nutritional Immunology at the University of Southampton.
She has researched the influences of early nutrition and the maternal diet in pregnancy on offspring
health and development. Her early research demonstrated that the neonatal immune responses of
babies who subsequently developed allergic symptoms were different from babies who did not.
Dr. Miles’ research has also showed that the supplementation of healthy humans with omega-3
polyunsaturated fatty acids results in a dose-dependent increase in omega-3 fatty acids in plasma and
immune cells and decreases ex vivo immune and inflammatory mediator production. Furthermore,
increasing the intake of omega-3 fatty acids in pregnancy in families with an increased risk of allergic
disease results in a decrease in ex vivo production of mediators involved in the allergic response.
Dr Miles has published over 76 peer-reviewed research papers, 4 book chapters and 17 reviews based
on her research interests.
Caroline E Childs is Lecturer in Nutritional Sciences at the University of Southampton. Her research
to date has focused on nutrients such as dietary fatty acids, probiotics and prebiotics and has assessed
the effect of nutrient interventions on outcomes such as tissue composition, immune function,
inflammatory status, immunosenescence and the gut microbiota. In 2010, she was named as one
of the top three young investigators in the category “Lipids and Nutrition” at the International
Society for the Study of Fatty Acids and Lipids conference. Dr Childs is Co-Chair of ILSI Europe’s
Nutrition, Immunity and Inflammation Task Force and a member of the ILSI Europe Expert Group
on Determinants of Immune Competence. Dr Childs is a Nutrition Society Ambassador for the
University of Southampton and Regional Representative for the Association for Nutrition. She is
an editorial board member of The Journal of Nutrition and sits on the Editorial Advisory Board of
Nutrition Bulletin.
vii
nutrients
Editorial
Diet and Immune Function
Caroline E. Childs 1 , Philip C. Calder 1,2 and Elizabeth A. Miles 1, *
1 Human Development and Health, Faculty of Medicine, University of Southampton,
Southampton SO16 6YD, UK
2 NIHR Southampton Biomedical Research Centre, University Hospital Southampton NHS Foundation Trust
and University of Southampton, Southampton SO16 6YD, UK
* Correspondence: e.a.miles@soton.ac.uk; Tel.: +44(0)23-8120-6925
Abstract: A well-functioning immune system is critical for survival. The immune system must be
constantly alert, monitoring for signs of invasion or danger. Cells of the immune system must be able
to distinguish self from non-self and furthermore discriminate between non-self molecules which are
harmful (e.g., those from pathogens) and innocuous non-self molecules (e.g., from food). This Special
Issue of Nutrients explores the relationship between diet and nutrients and immune function. In this
preface, we outline the key functions of the immune system, and how it interacts with nutrients
across the life course, highlighting the work included within this Special Issue. This includes the
role of macronutrients, micronutrients, and the gut microbiome in mediating immunological effects.
Nutritional modulation of the immune system has applications within the clinical setting, but can
also have a role in healthy populations, acting to reduce or delay the onset of immune-mediated
chronic diseases. Ongoing research in this field will ultimately lead to a better understanding of
the role of diet and nutrients in immune function and will facilitate the use of bespoke nutrition to
improve human health.
(Treg), which are CD4-bearing T cells vital in maintaining immune tolerance to allow the immune
system to ignore non-harmful non-self (such as food, pollen, and environmental antigens such as
latex). Thus, the role of T cells is coordinating an appropriate immune response following immune
stimulation or challenge.
The other lymphocytes of the adaptive immune system are the B cells, which are responsible for
antibody or immunoglobulin (Ig) production. Like T cells, B cells respond specifically to an antigen.
They can differentiate into short-lived plasma cells, which produce Igs in the short term, or can become
long-lived plasma cells. Igs are pathogen-specific molecules, which help the immune system to recognise
and destroy pathogens. The B cells can differentiate into plasma cells, which produce one of five classes of
Ig (IgM, IgD, IgG, IgA, and IgE). Each class of Ig has a specialised role [3]. IgM is the first Ig expressed
during development, is often found as a multimeric molecule (e.g., pentameric), and can bind an antigen
to identify it for destruction by immune cells. IgD is found in low concentrations in the plasma and the
specialist role of IgD is not yet clear. IgG is the predominant Ig class and can persist for long periods.
It has important roles in antigen labelling, resulting in more effective removal. IgA can be found in the
serum (mostly as a monomer) and at mucosal surfaces (normally as a dimer). At the mucosal surface,
IgA protects against bacteria and or viruses, preventing infection. IgA also has an important role in
neutralising food antigens and helping to maintain immune tolerance to food antigens (preventing the
development of food allergy) [4]. IgE has a role in clearance of extracellular parasites (e.g., helminths) but
when produced inappropriately to innocuous environmental and food antigens, has an important role in
IgE-mediated allergy. B cells go through a process called class switching to set the class of Ig that the
plasma cells derived from them will produce. B cell class switching is controlled by the cytokines present,
particularly IL-4, IL-6, and IFN-γ secreted from Th cells [5].
T and B cells can specialise to become memory cells, which persist permanently or for very long
periods and are able to recognise the antigen if encountered again and elicit a rapid, pathogen-specific
immune response.
The effective deployment of the immune system against pathogens or harmful signals and the
swift resolution of the immune response is required for survival. The fighting of infection is only
one piece of the puzzle. A fulminating immune response is costly in terms of energy expended and
results in damage to the host tissues; thus, rapid and complete resolution of an immune response is
also key. Cytokines play a role in resolution of immune responses. IL-10, which is produced by a range
of immune cells including Tregs, has anti-inflammatory actions including suppressing inflammatory
cytokine production [6].
The instigation of an immune response and the activities of the immune cells results in inflammation
(seen as redness, swelling, and the feeling of heat and pain), which are signs of the damage to the
tissue going on whilst the immune system does its work. This is an expected outcome of an effective
immune response. Increasingly there is concern that modern lifestyle changes have resulted in the
promotion of ongoing, low-grade, whole-body (systemic) inflammation caused by immune and other
cells (e.g., adipocytes, the cells that store lipids in fat tissue) [7]. Such exposures may include diet
quality and quantity [8].
2
Nutrients 2019, 11, 1933
immune system throughout the life course or in reducing chronic inflammation. For example, the amino
acid arginine is essential for the generation of nitric oxide by macrophages, and the micronutrients
vitamin A and zinc regulate cell division and so are essential for a successful proliferative response
within the immune system.
Undernutrition is well understood to impair immune function, whether as a result of food
shortages or famines in developing countries, or as a result of malnutrition arising from periods
of hospitalisation in developed countries. The extent of impairment that results will depend upon
the severity of the deficiency, whether there are nutrient interactions to consider, the presence of
infection, and the age of the subject [9]. A single nutrient can also exert multiple diverse immunological
effects, such as in the case of vitamin E, where it has a role as both antioxidant, inhibitor of protein
kinase C activity, and potentially interacting with enzymes and transport proteins [10]. For some
micronutrients, excessive intake can also be associated with impaired immune responses. For example,
supplementation with iron can increase morbidity and mortality of those in malaria endemic regions.
As well as nutrition having the potential to effectively treat immune deficiencies related to poor
intake, there is a great deal of research interest in whether specific nutrient interventions can further
enhance immune function in sub-clinical situations, and so prevent the onset of infections or chronic
inflammatory diseases.
3
Nutrients 2019, 11, 1933
IL-10 [18]. Prebiotics can also enhance barrier function, in addition to their role as substrates for
bacterial metabolism [19]. Santiago-Lopez et al. have investigated the effect of fermented milk on a
murine model of inflammatory bowel disease [18] and demonstrated a reduction in serum IL-17 and
IFN-γ following fermented milk consumption when compared with the control group.
4
Nutrients 2019, 11, 1933
and other co-morbidities (e.g., obesity, cardiovascular disease, insulin resistance). Interestingly, in a
study in healthy adults, increasing age was found to be a risk factor for chronic systemic inflammation,
independent of other risk factors such as body mass index, blood pressure, and blood lipid profiles [30].
The rising worldwide prevalence of obesity in children and adults is of grave concern. Obesity
and over nutrition are strongly associated with chronic inflammation, metabolic perturbation, and
higher risk for a number of chronic diseases including cardiovascular disease, stroke, type 2 diabetes
mellitus, and chronic liver disease. This metabolism-induced inflammation associated with obesity
is termed metaflammation, and the Western diet is a known risk factor [31,32]. The Western diet is
characterised by a diet high in sugar, trans and saturated fats, but low in complex carbohydrates, fibre,
micronutrients, and other bioactive molecules such as polyphenols and omega 3 polyunsaturated fatty
acids. The mechanisms by which the Western diet predispose individuals to metaflammation are still
under investigation. However, one mechanism which has been reported is the increased uptake of
lipopolysaccharide (LPS, a constituent of gram-negative bacterial cells walls), from microbes in the gut
because of increased gut leakiness. This LPS is sensed by cells of the innate immune system through
toll-like receptor 4 (TLR4). Activation of TLR4 by LPS will induce an inflammatory response by the
immune cells. Certain nutrients, notably long-chain omega 3 polyunsaturated fatty acids, can interfere
with TLR4 activation and, thus, can ameliorate this inflammatory signal. Rogero et al. describe the
relationship between obesity and inflammation and explores the immune pathway for this mechanism
and the anti-inflammatory roles of omega 3 fatty acids in this process [33].
Interestingly, in juxtaposition with the review by Rogero et al. on inflammation in obesity,
Dalton and colleagues report a study into systemic inflammation in individuals with the serious
psychological eating disorder, anorexia nervosa [34]. They show that in a severely undernourished
state, there are indications of systemic inflammation with an increased serum concentration of IL-6
when compared with healthy control participants. IL-6 is a classically inflammatory cytokine produced
by immune and other cells. Whether this inflammation is the result of the impact of undernutrition
or whether the clinical condition is the result of pre-existing inflammation is a matter that remains
to be determined. It has been shown that patients with clinical depression have increased systemic
inflammation suggesting that inflammation may have a bearing on mental health and wellbeing [35].
In contrast with the Western diet, the Mediterranean diet is rich in vegetables, fruit, nuts, legumes,
fish, and ‘healthy’ dietary fats. The Mediterranean diet is associated with a reduced risk of chronic
disease such as cardiovascular disease, cancer, and more recently Alzheimer’s disease [36]. A range of
bioactive compounds found in fruits and vegetables have been reported to offer one explanation for the
protective effect of diets rich in fruits and vegetables (e.g., Mediterranean diet) on the reduction of risk
for developing non-communicable diseases attributed to chronic inflammation (e.g., cardiovascular
disease). One family of molecules, which are known to have a role in regulation of inflammation
are the dietary polyphenols [37]. Yahfoufi et al. explain the mechanisms by which polyphenols
can be immunomodulatory and anti-inflammatory and explore the evidence for the role of dietary
polyphenols in reducing the risk of cardiovascular disease, some neurological diseases, and cancer [38].
5
Nutrients 2019, 11, 1933
Selenium is a trace element that, like zinc, has critical functional, structural, and enzymatic roles,
in a range of proteins. Poor selenium status is associated with a higher risk for range of chronic diseases
including cancer and cardiovascular disease [43]. In addition to critical roles in many non-immune
tissues within the body, selenium is important for optimal immune function. Avery and Hoffman
explain the role of selenium in immunobiology and the mechanisms by which selenoproteins regulate
immunity. The evidence for the significance of selenium status in infectious diseases including human
immunodeficiency virus infection is reviewed [44].
Glutamine is a nonessential amino acid that provides an important energy source for many cell types
including those involved in immune responses. It also serves as a precursor for nucleotide synthesis,
particularly relevant for rapidly dividing cells such as the immune cells during an immune response.
During infection, the rate of glutamine consumption by immune cells is equivalent or greater than that
for glucose. Glutamine has roles in the functions of a number of immune cells including neutrophils,
macrophages, and lymphocytes [45]. In catabolic conditions (e.g., infection, inflammation, trauma),
glutamine is released into the circulation, an essential process controlled by metabolic organs such as the
liver, gut, and skeletal muscles. Despite this adaptation, a significant depletion of glutamine is seen in the
plasma and tissues in critical illness, which has provided a rationale for the use of in clinical nutrition
supplementation of critically ill patients. How glutamine homeostasis is maintained and when and how
to utilise glutamine in the clinical setting is explored in a review by Cruzat et al. [45].
The vitamin D receptor (VDR) is a nuclear receptor that can directly affect gene expression [46].
The presence of VDR in the majority of immune cells immediately suggests an important role
for this micronutrient in immune cell activities [47]. Furthermore, vitamin D-activating enzyme
1-α-hydroxylase (CYP27B1), which results in the active metabolite 1 α,25-dihydroxyvitamin D3
(1,25(OH)2 D3 ), is expressed in many types of immune cells. Ligation of VDR by 1,25(OH)2 D3 can elicit
the production of antimicrobial proteins and influence cytokine production by immune cells [47,48].
Sassi, Tamone, and d’Amelio have reviewed the evidence for the role of the nutrient vitamin D in the
innate and adaptive immune systems [16].
7. Conclusions
In this Special Issue of Nutrients, the collected works provide a breadth of reviews and research
indicating the key influence of nutrients and nutrition on immune responses in health and disease and
across the life course. Nutrients may impact directly or indirectly upon immune cells causing changes
in their function or may exert effects via changes in the gut microbiome. A better understanding
of the role of nutrients in immune function will facilitate the use of bespoke nutrition to improve
human health.
Author Contributions: Conceptualization, E.A.M.; writing—original draft preparation, E.A.M. and C.E.C.;
writing—review and editing, E.A.M., P.C.C., and C.E.C.
Funding: This article received no external funding.
Acknowledgments: This article received no specific grant from any funding agency, commercial or
not-for-profit sectors.
Conflicts of Interest: C.E.C. is member of the ILSI Europe Expert Group on Determinants of Immune Competence
and Co-Chair of ILSI Europe’s Nutrition, Immunity and Inflammation Task Force. C.E.C. receives research funding
from HOST Therabiomics and honoraria to speak at an event organised by Yakult. P.C.C. has research funding
from Bayer, has received research study products from Christian Hansen, and acts as a consultant/adviser to BASF,
DSM, Cargill, Smartfish and Pfizer. E.A.M. has no conflicts of interest to declare.
References
1. Romagnani, S. T-cell subsets (Th1 versus Th2). Ann. Allergy Asthma Immunol. 2000, 85, 9–18. [CrossRef]
2. Zhu, J.; Yamane, H.; Paul, W.E. Differentiation of effector CD4 T cell populations. Annu. Rev. Immunol.
2010, 28, 445–489. [CrossRef] [PubMed]
6
Nutrients 2019, 11, 1933
3. Schroeder, H.W., Jr.; Cavacini, L. Structure and function of immunoglobulins. J. Allergy Clin. Immunol.
2010, 125, 41–52. [CrossRef]
4. Berin, M.C. Mucosal antibodies in the regulation of tolerance and allergy to foods. Semin. Immunopathol.
2012, 34, 633–642. [CrossRef] [PubMed]
5. Vazquez, M.I.; Catalan-Dibene, J.; Zlotnik, A. B cells responses and cytokine production are regulated by
their immune microenvironment. Cytokine 2015, 74, 318–326. [CrossRef]
6. Saraiva, M.; O’Garra, A. The regulation of IL-10 production by immune cells. Nat. Rev. Immunol.
2010, 10, 170–181. [CrossRef] [PubMed]
7. Calder, P.C.; Ahluwalia, N.; Brouns, F.; Buetler, T.; Clement, K.; Cunningham, K.; Esposito, K.; Jonsson, L.S.;
Kolb, H.; Lansink, M.; et al. Dietary factors and low-grade inflammation in relation to overweight and
obesity. Br. J. Nutr. 2011, 106, 5–78. [CrossRef]
8. Calder, P.C.; Bosco, N.; Bourdet-Sicard, R.; Capuron, L.; Delzenne, N.; Dore, J.; Franceschi, C.; Lehtinen, M.J.;
Recker, T.; Salvioli, S.; et al. Health relevance of the modification of low grade inflammation in ageing
(inflammageing) and the role of nutrition. Ageing Res. Rev. 2017, 40, 95–119. [CrossRef]
9. Calder, P.C.; Jackson, A.A. Undernutrition, infection and immune function. Nutr. Res. Rev. 2000, 13, 3–29.
[CrossRef]
10. Lee, G.Y.; Han, S.N. The Role of Vitamin E in Immunity. Nutrients 2018, 10, 1614. [CrossRef]
11. Macdonald, T.T.; Monteleone, G. Immunity, inflammation, and allergy in the gut. Science 2005, 307, 1920–1925.
[CrossRef] [PubMed]
12. Hill, C.; Guarner, F.; Reid, G.; Gibson, G.R.; Merenstein, D.J.; Pot, B.; Morelli, L.; Canani, R.B.; Flint, H.J.;
Salminen, S.; et al. Expert consensus document. The International Scientific Association for Probiotics and
Prebiotics consensus statement on the scope and appropriate use of the term probiotic. Nat. Rev. Gastroenterol.
Hepatol. 2014, 11, 506–514. [CrossRef] [PubMed]
13. Gibson, G.R.; Hutkins, R.; Sanders, M.E.; Prescott, S.L.; Reimer, R.A.; Salminen, S.J.; Scott, K.; Stanton, C.;
Swanson, K.S.; Cani, P.D.; et al. Expert consensus document: The International Scientific Association for
Probiotics and Prebiotics (ISAPP) consensus statement on the definition and scope of prebiotics. Nat. Rev.
Gastroenterol. Hepatol. 2017, 14, 491–502. [CrossRef] [PubMed]
14. Hansen, N.W.; Sams, A. The Microbiotic Highway to Health-New Perspective on Food Structure, Gut
Microbiota, and Host Inflammation. Nutrients 2018, 10, 1590. [CrossRef] [PubMed]
15. Bischoff, S.C.; Barbara, G.; Buurman, W.; Ockhuizen, T.; Schulzke, J.D.; Serino, M.; Tilg, H.; Watson, A.;
Wells, J.M. Intestinal permeability—A new target for disease prevention and therapy. BMC Gastroenterol.
2014, 14, 189. [CrossRef] [PubMed]
16. Sassi, F.; Tamone, C.; D’Amelio, P. Vitamin D: Nutrient, Hormone, and Immunomodulator. Nutrients
2018, 10, 1656. [CrossRef] [PubMed]
17. Kiewiet, M.B.G.; Faas, M.M.; de Vos, P. Immunomodulatory Protein Hydrolysates and Their Application.
Nutrients 2018, 10, 904. [CrossRef]
18. Santiago-Lopez, L.; Hernandez-Mendoza, A.; Mata-Haro, V.; Vallejo-Cordoba, B.; Wall-Medrano, A.;
Astiazaran-Garcia, H.; Estrada-Montoya, M.D.C.; Gonzalez-Cordova, A.F. Effect of Milk Fermented with
Lactobacillus fermentum on the Inflammatory Response in Mice. Nutrients 2018, 10, 1039. [CrossRef]
19. McKeen, S.; Young, W.; Mullaney, J.; Fraser, K.; McNabb, W.C.; Roy, N.C. Infant Complementary Feeding of
Prebiotics for theMicrobiome and Immunity. Nutrients 2019, 11, 364. [CrossRef]
20. Kamemura, N.; Tada, H.; Shimojo, N.; Morita, Y.; Kohno, Y.; Ichioka, T.; Suzuki, K.; Kubota, K.; Hiyoshi, M.;
Kido, H. Intrauterine sensitization of allergen-specific IgE analyzed by a highly sensitive new allergen
microarray. J. Allergy Clin. Immunol. 2012, 130, 113–121. [CrossRef]
21. Donovan, S.M.; Comstock, S.S. Human Milk Oligosaccharides Influence Neonatal Mucosal and Systemic
Immunity. Ann. Nutr. Metab. 2016, 69, 42–51. [CrossRef] [PubMed]
22. Plaza-Diaz, J.; Fontana, L.; Gil, A. Human Milk Oligosaccharides and Immune System Development.
Nutrients 2018, 10, 1038. [CrossRef] [PubMed]
23. Torres-Castro, P.; Abril-Gil, M.; Rodriguez-Lagunas, M.J.; Castell, M.; Perez-Cano, F.J.; Franch, A. TGF-beta2,
EGF, and FGF21 Growth Factors Present in Breast Milk Promote Mesenteric Lymph Node Lymphocytes
Maturation in Suckling Rats. Nutrients 2018, 10, 1171. [CrossRef] [PubMed]
7
Nutrients 2019, 11, 1933
24. Kim, H.; Sitarik, A.R.; Woodcroft, K.; Johnson, C.C.; Zoratti, E. Birth Mode, Breastfeeding, Pet Exposure, and
Antibiotic Use: Associations With the Gut Microbiome and Sensitization in Children. Curr. Allergy Asthma
Rep. 2019, 19, 22. [CrossRef] [PubMed]
25. Kuper, C.F.; van Bilsen, J.; Cnossen, H.; Houben, G.; Garthoff, J.; Wolterbeek, A. Development of immune
organs and functioning in humans and test animals: Implications for immune intervention studies.
Reprod. Toxicol. 2016, 64, 180–190. [CrossRef] [PubMed]
26. Crooke, S.N.; Ovsyannikova, I.G.; Poland, G.A.; Kennedy, R.B. Immunosenescence: A systems-level overview
of immune cell biology and strategies for improving vaccine responses. Exp. Gerontol. 2019, 124, 110632.
[CrossRef] [PubMed]
27. Berzins, S.P.; Uldrich, A.P.; Sutherland, J.S.; Gill, J.; Miller, J.F.; Godfrey, D.I.; Boyd, R.L. Thymic regeneration:
Teaching an old immune system new tricks. Trends Mol. Med. 2002, 8, 469–476. [CrossRef]
28. Salanitro, A.H.; Ritchie, C.S.; Hovater, M.; Roth, D.L.; Sawyer, P.; Locher, J.L.; Bodner, E.; Brown, C.J.;
Allman, R.M. Inflammatory biomarkers as predictors of hospitalization and death in community-dwelling
older adults. Arch. Gerontol. Geriatr. 2012, 54, 387–391. [CrossRef]
29. Maggini, S.; Pierre, A.; Calder, P.C. Immune Function and Micronutrient Requirements Change over the Life
Course. Nutrients 2018, 10, 1531. [CrossRef]
30. Miles, E.A.; Rees, D.; Banerjee, T.; Cazzola, R.; Lewis, S.; Wood, R.; Oates, R.; Tallant, A.; Cestaro, B.; Yaqoob, P.;
et al. Age-related increases in circulating inflammatory markers in men are independent of BMI, blood
pressure and blood lipid concentrations. Atherosclerosis 2008, 196, 298–305. [CrossRef]
31. Hotamisligil, G.S. Inflammation, metaflammation and immunometabolic disorders. Nature 2017, 542, 177–185.
[CrossRef] [PubMed]
32. Christ, A.; Latz, E. The Western lifestyle has lasting effects on metaflammation. Nat. Rev. Immunol. 2019, 19, 267–268.
[CrossRef] [PubMed]
33. Rogero, M.M.; Calder, P.C. Obesity, Inflammation, Toll-Like Receptor 4 and Fatty Acids. Nutrients 2018, 10, 432.
[CrossRef] [PubMed]
34. Dalton, B.; Campbell, I.C.; Chung, R.; Breen, G.; Schmidt, U.; Himmerich, H. Inflammatory Markers in
Anorexia Nervosa: An Exploratory Study. Nutrients 2018, 10, 1573. [CrossRef] [PubMed]
35. Dantzer, R.; O’Connor, J.C.; Freund, G.G.; Johnson, R.W.; Kelley, K.W. From inflammation to sickness and
depression: When the immune system subjugates the brain. Nat. Rev. Neurosci. 2008, 9, 46–56. [CrossRef]
[PubMed]
36. Dinu, M.; Pagliai, G.; Casini, A.; Sofi, F. Mediterranean diet and multiple health outcomes: An umbrella
review of meta-analyses of observational studies and randomised trials. Eur. J. Clin. Nutr. 2018, 72, 30–43.
[CrossRef] [PubMed]
37. Rahman, I.; Biswas, S.K.; Kirkham, P.A. Regulation of inflammation and redox signaling by dietary
polyphenols. Biochem. Pharmacol. 2006, 72, 1439–1452. [CrossRef]
38. Yahfoufi, N.; Alsadi, N.; Jambi, M.; Matar, C. The Immunomodulatory and Anti-Inflammatory Role of
Polyphenols. Nutrients 2018, 10, 1618. [CrossRef]
39. World Health Organization Sepsis. Available online: https://www.who.int/news-room/fact-sheets/detail/
sepsis (accessed on 26 July 2019).
40. Alker, W.; Haase, H. Zinc and Sepsis. Nutrients 2018, 10, 976. [CrossRef]
41. Andreini, C.; Banci, L.; Bertini, I.; Rosato, A. Counting the zinc-proteins encoded in the human genome.
J. Proteome Res. 2006, 5, 196–201. [CrossRef]
42. Ibs, K.H.; Rink, L. Zinc-altered immune function. J. Nutr. 2003, 133, 1452–1456. [CrossRef] [PubMed]
43. Rayman, M.P. Selenium and human health. Lancet 2012, 379, 1256–1268. [CrossRef]
44. Avery, J.C.; Hoffmann, P.R. Selenium, Selenoproteins, and Immunity. Nutrients 2018, 10, 1203. [CrossRef]
[PubMed]
45. Cruzat, V.; Macedo Rogero, M.; Noel Keane, K.; Curi, R.; Newsholme, P. Glutamine: Metabolism and Immune
Function, Supplementation and Clinical Translation. Nutrients 2018, 10, 1654. [CrossRef] [PubMed]
46. Haussler, M.R.; Whitfield, G.K.; Haussler, C.A.; Hsieh, J.C.; Thompson, P.D.; Selznick, S.H.; Dominguez, C.E.;
Jurutka, P.W. The nuclear vitamin D receptor: Biological and molecular regulatory properties revealed.
J. Bone Miner. Res. Off. J. Am. Soc. Bone Miner. Res. 1998, 13, 325–349. [CrossRef] [PubMed]
8
Nutrients 2019, 11, 1933
47. Baeke, F.; Takiishi, T.; Korf, H.; Gysemans, C.; Mathieu, C. Vitamin D: Modulator of the immune system.
Curr. Opin. Pharmacol. 2010, 10, 482–496. [CrossRef] [PubMed]
48. Wang, T.T.; Nestel, F.P.; Bourdeau, V.; Nagai, Y.; Wang, Q.; Liao, J.; Tavera-Mendoza, L.; Lin, R.; Hanrahan, J.W.;
Mader, S.; et al. Cutting edge: 1,25-dihydroxyvitamin D3 is a direct inducer of antimicrobial peptide gene
expression. J. Immunol. 2004, 173, 2909–2912. [CrossRef] [PubMed]
© 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
9
nutrients
Article
Inflammatory Markers in Anorexia Nervosa:
An Exploratory Study
Bethan Dalton 1, *, Iain C. Campbell 1 , Raymond Chung 2,3 , Gerome Breen 2,3 , Ulrike Schmidt 1,4
and Hubertus Himmerich 1,4
1 Section of Eating Disorders, Department of Psychological Medicine, Institute of Psychiatry,
Psychology & Neuroscience, King’s College London, London SE5 8AF, UK; iain.campbell@kcl.ac.uk (I.C.C.);
ulrike.schmidt@kcl.ac.uk (U.S.); hubertus.himmerich@kcl.ac.uk (H.H.)
2 MRC Social, Genetic, and Developmental Psychiatry Centre, Institute of Psychiatry, Psychology &
Neuroscience, King’s College London, London SE5 8AF, UK; raymond.chung@kcl.ac.uk (R.C.);
gerome.breen@kcl.ac.uk (G.B.)
3 National Institute for Health Research Biomedical Research Centre for Mental Health, Institute of Psychiatry,
Psychology & Neuroscience at the Maudsley Hospital and King’s College London, London SE5 8AF, UK
4 South London and Maudsley NHS Foundation Trust, Bethlem Royal Hospital, Monks Orchard Road,
Beckenham, Kent BR3 3BX, UK
* Correspondence: bethan.l.dalton@kcl.ac.uk; Tel.: +44-0207-848-0183
Abstract: Inflammation has been suggested to play a pathophysiological role in anorexia nervosa
(AN). In this exploratory cross-sectional study, we measured serum concentrations of 40 inflammatory
markers (including cytokines, chemokines, and adhesion molecules) and brain-derived neurotrophic
factor (BDNF) in people with AN (n = 27) and healthy controls (HCs) (n = 13). Many of these
inflammatory markers had not been previously quantified in people with AN. Eating disorder
(ED) and general psychopathology symptoms were assessed. Body mass index (BMI) and body
composition data were obtained. Interleukin (IL)-6, IL-15, and vascular cell adhesion molecule
(VCAM)-1 concentrations were significantly elevated and concentrations of BDNF, tumor necrosis
factor (TNF)-β, and vascular endothelial growth factor (VEGF)-A were significantly lower in AN
participants compared to HCs. Age, BMI, and percentage body fat mass were identified as potential
confounding variables for several of these inflammatory markers. Of particular interest is that most of
the quantified markers were unchanged in people with AN, despite them being severely underweight
with evident body fat loss, and having clinically significant ED symptoms and severe depression and
anxiety symptoms. Future research should examine the replicability of our findings and consider the
effect of additional potential confounding variables, such as smoking and physical activity, on the
relationship between AN and inflammation.
1. Introduction
Anorexia nervosa (AN) is a serious psychiatric disorder characterised by low body weight due to
food restriction and weight-control behaviours, such as excessive exercise and self-induced vomiting,
together with an intense fear of weight gain and disturbed body perception [1]. Altered concentrations
of inflammatory markers, in particular cytokines, have been reported in people with AN [2,3].
Cytokines are cell signalling molecules produced by a range of cells (e.g., microglia, astrocytes) in the
brain and the periphery (e.g., by macrophages and T-lymphocytes) and are essential in coordinating
responses to infection [4]. In addition, changes in the circulating concentrations and production of
cytokines have been associated with a range of disease states, including obesity [5] and diabetes [6],
as well as depression [7], schizophrenia [8], and eating disorders (EDs) [2,3].
Research in AN has primarily focused on pro-inflammatory cytokines, which promote and up-regulate
inflammatory reactions [4]. Recent meta-analyses have concluded that the pro-inflammatory cytokines
tumor necrosis factor (TNF)-α and interleukin (IL)-6, are elevated in people with AN, compared to
healthy individuals (for reviews see: [2,3]). However, few studies have quantified the concentrations
of cytokines in other categories, such as T-helper (TH )-1, TH 2, and anti-inflammatory cytokines (e.g.,
IL-10), the latter of which play an immunomodulatory role by reducing inflammation [9]. An example
of one such cytokine yet to be measured in people with AN is TNF-β, which is produced by TH 1
cells. TNF-β performs a variety of important roles in immune regulation [10,11], but has also been
implicated in the regulation of the commensal gut microbiota [12–14], which appears to be involved
in the pathology of AN [15–17]. Additionally, a number of cytokines implicated in other disorders,
such as depression and obesity, are yet to be measured in AN. One example is IL-17, a TH 17 cytokine
that has been reported to predict treatment response in people with depression [18], and seems to
be involved in the pathophysiology of schizophrenia [19] and the molecular and cellular effects of
antipsychotics [20].
Chemokines are a subcategory of smaller cytokines known to induce chemotaxis, with some
also having a homeostatic function in relation to haematopoiesis, immune surveillance, and adaptive
immune system responses [21,22]. The chemokines RANTES, monocyte chemoattractant protein
(MCP)-1, and fractalkine have been measured in two studies in people with AN [23,24]. Similarly,
adhesion molecules, which mediate the binding of cells in the immune system [25], have been measured
in one study in a sample of people with AN [26]. Circulating concentrations of vascular cell adhesion
molecule (VCAM)-1 have been reported to be elevated in people with AN compared to healthy
participants, but intercellular adhesion molecule (ICAM)-1 did not differ between the groups.
Cytokines and chemokines impact several biological domains implicated in the pathophysiology
of AN, including the modulation of neurotransmitter systems, neuroendocrine functioning, and neural
plasticity [27–31]. For example, in the depression literature, it has been hypothesised that elevated
pro-inflammatory cytokine levels may lead to symptoms of depression, partly via their disruption of
growth factor production, e.g., brain-derived neurotropic factor (BDNF) [32] and vascular endothelial
growth factor (VEGF)-A [33], which has a subsequent effect on adult neurogenesis [28,34]. Disruption
to these biological processes can then lead to alterations in mental state, including affect, learning and
memory, and behaviour (e.g., depressive-like behaviours) [28,35].
A number of factors, including body mass index (BMI), age, medication, and smoking status,
have been reported to influence cytokine concentrations [36]. These may be potential confounding
factors in studies of the role of cytokines in AN, particularly given the low weight seen in AN,
the tendency for research in EDs to focus on adolescents and young adults [2], and research indicating
that people with EDs report higher rates of smoking than healthy controls [37]. Previous studies
have not assessed the potential impact of depression symptoms on cytokine concentrations in AN.
A pro-inflammatory profile has been identified in people with depression [38] and the comorbidity
between AN and unipolar depression is of significant clinical relevance, as approximately 40% of
people receiving treatment for AN also suffer from depression [39]. Therefore, it is unclear as to
whether the alterations observed in cytokine concentrations are due to the AN or symptoms of
comorbid disorders, such as depression.
Few studies have considered a broad range of cytokines and other markers involved in
inflammatory processes and their potential role in the biological profile of AN. Therefore, in this
exploratory cross-sectional study, we measured a variety of inflammatory markers in a sample of AN
participants and healthy controls (HCs) to determine whether these markers are altered in AN. Several
of these inflammatory markers have not been previously quantified in people with AN. A secondary
objective was to test for the effects of potential confounders on concentrations of the inflammatory
11
Nutrients 2018, 10, 1573
markers, including age and BMI, and explore the effect of current symptom severity on markers of
inflammation in people with AN.
2.2. Measures
2.2.2. Anthropometry
Height and body weight were measured, and from these measurements, BMI (kg/m2 ) was
calculated. Body composition was also measured using a portable and non-invasive Inbody S10
machine (Inbody Co., Ltd., Seoul, Korea) which uses the Bioelectrical Impedance Analysis (BIA)
measurement method. Following the input of height and weight details, this machine provides data
on muscle and fat, bone mineral content, intracellular and extracellular water, protein, and minerals.
The calculations used to do this are based on the assumption that the body is a cylindrical-shaped
conductor. Resistance is low in lean tissue (as it contains the majority of intracellular and extracellular
fluid and is thus a good conductor of electrical current), and fat mass is high in resistance as it is does
not contain any water (and thus does not conduct electrical current). Based on the assumption that
impedance (resistance) is proportional to total body water, predictive equations then determine total
body water, total body fat, and lean tissue mass. Given that adipose tissue has been implicated in the
genesis of cytokines and produces certain pro-inflammatory cytokines (e.g., IL-6), we focused on the
association between inflammatory markers and body fat percentage and did not include other body
composition parameters in our analyses.
12
Nutrients 2018, 10, 1573
A global score can be calculated, and items can also be categorised and scored into the following four
subscales: restraint, eating concern, weight concern, and shape concern. Related psychopathology
was assessed using the Depression Anxiety Stress Scale 21-Version (DASS-21) [44]. This is a 21-item
questionnaire measuring symptoms of general psychopathology over the previous seven days. As well
as a total score, a score for the three subscales—depression, anxiety, and stress—can be calculated.
Additional measures related to physical activity were also collected and findings are reported
elsewhere [40,41].
13
Nutrients 2018, 10, 1573
illness duration or ED symptoms, as measured by the EDE-Q, as the independent variable. To test
the effect of general psychopathology on inflammatory marker concentrations, linear regressions
in the AN participants, with the log-transformed inflammatory marker as the dependent variable
and the total DASS-21 score as the independent variable, were conducted. For both sets of analyses,
studentized residuals greater than ±3 standard deviations were deemed to be outliers and were
removed, and assumptions were tested and met.
The level of significance was set at p < 0.05, and as this was an exploratory study, levels of
significance were not adjusted for multiple testing.
3. Results
Demographic, anthropometric, and clinical characteristics of the AN participants and HCs are
presented in Table 1. All participants were female. Mean age did not significantly differ between the
AN and HC groups (U = 144, z = −1.36, p = 0.1735). Seven participants with AN reported being a
current smoker, with an average of 9.14 ± 5.90 cigarettes smoked per day. As expected, AN participants
had lower BMI (t (38) = 7.88, p < 0.001) and percentage body fat (U = −22, z = 3.63, p = 0.0003) scores,
and higher EDE-Q scores (global score: U = 85.5, z = −4.87, p < 0.001) than HCs. The EDE-Q global
score for the AN participants was greater than the commonly used clinical cut-off score of 4 e.g., [46,47].
AN participants also reported greater depression, anxiety, and stress than HCs on the DASS-21 (total
score: U = 92.5, z = −4.67, p <0.001). Proposed cut-off scores [44] suggest that the level of severity that
AN participants reported was severe for depression, anxiety, and stress.
Table 1. Demographic, anthropometric, and clinical characteristics for AN participants and HCs.
SD—standard deviation; BMI—body mass index; AN-R—anorexia nervosa restricting type; AN-BP—anorexia
nervosa binge-eating/purging type; EDE-Q—Eating Disorder Examination—Questionnaire; DASS-21—Depression
Anxiety and Stress Scales-21 Version.
14
Table 2. Median serum concentrations (pg/mL), with interquartile range (IQR), minimum and maximum values, of inflammatory markers for HCs and
AN participants.
BDNF 13 17,375.24 7862.72 10,526.40 21,884.09 27 12,799.75 5947.87 7792.55 23,003.72 0.0315
TNF-β 13 0.86 0.32 0.43 53.22 27 0.60 0.19 0.16 1.22 0.0041
VEGF-A 13 471.96 220.76 205.98 749.41 27 288.82 155.41 133.70 973.73 0.0338
b. Concentrations significantly higher in AN compared to HCs
IL-6 13 0.38 0.29 0.10 1.11 27 0.49 0.90 0.20 3.35 0.0159
IL-15 13 2.54 0.39 1.91 3.18 27 2.90 0.81 1.57 5.24 0.0090
VCAM-1 13 612,378.30 92,337.87 402,297.20 739,682.80 27 709,059.60 220,832.10 434,057.50 1,069,997.00 0.0448
c. Concentrations not significantly different between groups
bFGF 13 11.01 10.73 2.17 21.13 27 12.13 13.39 1.15 65.25 0.3334
CRP 13 332,422.10 1,497,240.00 38,821.00 4,878,443.00 27 342,975.80 5,408,764.00 39,716.48 37,500,000.00 0.9424
Eotaxin 13 208.55 61.81 136.51 285.80 27 175.47 123.38 101.36 511.08 0.8511
Eotaxin-3 13 20.55 6.48 9.95 112.91 27 15.15 14.22 5.14 60.87 0.1058
15
Flt-1 13 61.83 11.64 27.93 86.21 27 68.69 28.37 11.99 117.06 0.1224
GM-CSF 13 0.15 0.21 0.02 0.88 22 0.19 0.17 0.01 0.35 0.8779
ICAM-1 13 658,988.00 183,427.10 517,283.10 861,717.20 27 701,224.80 244,044.70 415,274.60 977,044.00 0.1224
IFN-γ 13 3.94 1.64 2.11 6.84 27 4.64 5.72 2.11 64.28 0.1448
IL-1α 13 1.13 1.32 0.41 3.51 27 1.01 0.64 0.44 6.00 0.4105
IL-1β 9 0.20 0.33 0.04 1.86 21 0.19 0.19 0.01 4.07 0.7343
IL-2 6 0.12 0.24 0.03 0.43 13 0.15 0.16 0.01 0.85 0.9301
IL-4 13 0.08 0.07 0.01 0.20 25 0.05 0.05 0.00 5.24 0.1481
IL-5 12 1.09 0.87 0.43 3.45 24 1.29 1.01 0.07 4.39 0.7626
IL-7 13 13.66 6.11 8.76 20.13 27 12.58 6.80 5.82 26.20 0.8966
IL-8 13 33.51 62.10 8.44 503.23 27 23.09 92.94 5.80 388.41 0.4271
IL-10 13 0.18 0.10 0.06 0.41 27 0.28 0.26 0.04 4.40 0.3120
IL-12/IL-23p40 13 117.69 49.83 72.15 250.50 27 92.00 50.59 38.72 220.89 0.0806
IL-12p70 11 0.18 0.09 0.01 0.53 25 0.19 0.25 0.01 1.06 0.7572
IL-13 9 2.95 3.25 1.53 10.34 17 2.44 3.13 0.76 6.35 0.2692
IL-16 13 160.52 58.89 99.73 291.95 27 183.59 194.22 127.93 463.24 0.1702
Table 2. Cont.
MCP-1 13 208.55 126.78 140.94 406.88 27 191.53 77.31 121.51 345.28 0.3480
MCP-4 13 142.79 80.25 54.04 212.93 27 120.65 83.26 53.61 290.55 0.6133
MIP-1α 13 25.01 7.73 16.36 56.15 27 23.02 15.10 15.68 66.68 0.8624
MIP-1β 13 102.15 62.83 34.70 280.41 27 81.06 44.32 41.10 271.72 0.3334
PlGF 13 3.38 1.57 1.91 6.06 27 3.85 1.81 1.71 6.21 0.6754
SAA 13 2,683,488.00 1,218,569.00 318,141.50 13,200,000.00 27 2,681,116.00 9,488,702.00 525,220.00 164,000,000.00 0.8061
TARC 13 365.16 319.52 168.47 690.47 27 370.33 378.45 68.69 7680.34 0.7617
Tie-2 13 5847.41 2784.74 4065.41 8455.58 27 6212.88 3028.74 2440.01 10,077.48 0.7617
TNF-α 13 1.59 0.50 0.91 2.70 27 1.64 1.08 0.61 2.95 0.8966
VEGF-C 13 574.92 119.71 242.97 739.49 27 449.26 243.10 277.52 852.72 0.2919
VEGF-D 13 790.06 358.08 383.69 1126.91 27 757.22 413.20 458.68 1906.49 0.8061
Abbreviations: IQR—interquartile range; min—minimum; max—maximum; AN—anorexia nervosa; HC—healthy controls; BDNF—brain-derived neurotrophic factor; TNF—tumor
necrosis factor; VEGF—vascular endothelial growth factor; IL—interleukin; VCAM-1—vascular cell adhesion molecule-1; bFGF—basic fibroblast growth factor; CRP—C-reactive protein;
Flt-1—Fms-like tyrosine kinase-1; GM-CSF—granulocyte-macrophage colony-stimulating factor; ICAM-1—intercellular adhesion molecule-1, IFN-γ—interferon- γ; IP-10—interferon
16
γ-induced protein-10; MCP—monocyte chemoattractant protein; MIP—macrophage inflammatory protein; PlGF—placental growth factor; SAA—serum amyloid A; TARC—thymus and
activation-regulated chemokine; Tie-2—tyrosine kinase-2.
Nutrients 2018, 10, 1573
4. Discussion
17
Nutrients 2018, 10, 1573
18
Nutrients 2018, 10, 1573
therapeutic relevance. For example, we could consider the question as to whether antipsychotics,
such as olanzapine, which is approved for the treatment of schizophrenia and alters cytokine
production [65], might help patients with AN [31,66]. This may also be of particular interest in
AN as olanzapine has been shown to alter the gut microbiome, which could additionally contribute to
weight gain [67–70].
The findings of increased IL-6 and VCAM-1 serum concentrations in our sample of people with
AN compared to HCs are of less novelty, but indicate the reliability of our findings, as these results
have also been reported previously.
As described, there are already a number of studies that have consistently identified an association
between AN and the pro-inflammatory cytokine IL-6 [2]. IL-6 is an inducer of the acute-phase response,
which has been shown to have suppressive effects on food intake [71] and inhibit adipogenesis [72].
Our results replicated the findings of increased concentrations of IL-6 in people with AN, as compared
to HCs.
Víctor et al. [26] previously reported increased VCAM-1 serum levels in patients with AN.
VCAM-1 is a cell adhesion molecule with a key role in leukocyte recruitment from blood into tissue
and is thus important for cellular immune response [73]. Because of its wide distribution in human
tissues and organs, VCAM-1 has been implicated in the development of a variety of pathophysiological
states in the brain and in the body periphery, including autoimmune diseases, cardiovascular disease,
and infections [74].
It is unclear why these particular inflammatory markers are altered in people with AN compared
to HCs. However, there are a number of potential factors which may contribute to these alterations,
including stress and neuroendocrine functioning, genetics, the gut microbiota, early life stress,
and negative health behaviours (e.g., disturbed sleep, altered diet, smoking) [75].
19
Nutrients 2018, 10, 1573
5. Conclusions
This exploratory study measured a broad range of inflammatory markers, many of which had
not been previously assessed in AN. IL-15, VEGF-A, and TNF-β, for the first time, were shown to be
altered in people with AN in comparison to HCs. Previous findings regarding an elevation of IL-6 and
VCAM-1 and a reduction in BDNF in AN participants were replicated. We also considered age, BMI,
and percentage fat mass as potential confounding variables of concentrations of the inflammatory
markers. Our findings suggest that future research should include covariates in analyses of this
relationship to explore whether this may account for some of the group differences in inflammatory
markers observed in the current study. Finally, given that these inflammatory markers function as part
of a complex network, future studies in larger samples should consider developing a composite score
of cytokine concentrations.
20
Nutrients 2018, 10, 1573
Author Contributions: Conceptualization, G.B., U.S., and H.H.; Data curation, B.D.; Formal analysis, B.D.;
Funding acquisition, I.C.C., G.B., U.S., and H.H.; Investigation, B.D., I.C.C., R.C., and U.S.; Methodology, I.C.C.,
R.C., and U.S.; Project administration, U.S.; Supervision, I.C.C., G.B., U.S., and H.H.; Writing–original draft, B.D.;
Writing—review & editing, B.B., I.C.C., R.C., G.B., U.S., and H.H.
Funding: The ROSANA study was supported by an National Institute for Health Research (NIHR) Programme
Grant for Applied Research (RP-PG-0606-1043). The current study was funded by a studentship awarded to Bethan
Dalton by the Department of Psychological Medicine, King’s College London (KCL) and the Institute of Psychiatry,
Psychology and Neuroscience (IoPPN), KCL. Iain Campbell, Raymond Chung, Gerome Breen, and Ulrike Schmidt
receive salary support from the NIHR Mental Health Biomedical Research Centre at South London and Maudsley
NHS Foundation Trust and KCL. Ulrike Schmidt is supported by an NIHR Senior Investigator Award. The views
expressed are those of the author(s) and not necessarily those of the NHS, the NIHR, or the Department of Health
and Social Care. The APC was funded locally by monies supplied by Hubertus Himmerich.
Acknowledgments: We thank the participants of the ROSANA study and the research team who collected the
data between 2009 and 2013.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. American Psychiatric Association. Diagnostic and Statistical Manual of Mental Disorders, 5th ed.; American
Psychiatric Publishing: Arlington, VA, USA, 2013.
2. Dalton, B.; Bartholdy, S.; Robinson, L.; Solmi, M.; Ibrahim, M.A.A.; Breen, G.; Schmidt, U.; Himmerich, H.
A meta-analysis of cytokine concentrations in eating disorders. J. Psychiatr. Res. 2018, 103, 252–264. [CrossRef]
[PubMed]
3. Solmi, M.; Veronese, N.; Favaro, A.; Santonastaso, P.; Manzato, E.; Sergi, G.; Correll, C.U.
Inflammatory cytokines and anorexia nervosa: A meta-analysis of cross-sectional and longitudinal studies.
Psychoneuroendocrinology 2015, 51, 237–252. [CrossRef] [PubMed]
4. Dinarello, C.A. Proinflammatory cytokines. Chest 2000, 118, 503–508. [CrossRef] [PubMed]
5. Schmidt, F.M.; Weschenfelder, J.; Sander, C.; Minkwitz, J.; Thormann, J.; Chittka, T.; Mergl, R.; Kirkby, K.C.;
Fasshauer, M.; Stumvoll, M.; et al. Inflammatory cytokines in general and central obesity and modulating
effects of physical activity. PLoS ONE 2015, 10, e0121971. [CrossRef] [PubMed]
6. Xiao, J.; Li, J.; Cai, L.; Chakrabarti, S.; Li, X. Cytokines and diabetes research. J. Diabetes Res. 2014, 2014, 920613.
[CrossRef] [PubMed]
7. Lichtblau, N.; Schmidt, F.M.; Schumann, R.; Kirkby, K.C.; Himmerich, H. Cytokines as biomarkers in
depressive disorder: Current standing and prospects. Int. Rev. Psychiatry 2013, 25, 592–603. [CrossRef]
[PubMed]
8. Müller, N.; Weidinger, E.; Leitner, B.; Schwarz, M.J. The role of inflammation in schizophrenia. Front. Neurosci.
2015, 9, 372. [CrossRef] [PubMed]
9. Opal, S.M.; DePalo, V.A. Anti-inflammatory cytokines. Chest 2000, 117, 1162–1172. [CrossRef] [PubMed]
10. Ruddle, N.H. Lymphotoxin and TNF: How it all began-a tribute to the travelers. Cytokine Growth Factor Rev.
2014, 25, 83–89. [CrossRef] [PubMed]
11. Bauer, J.; Namineni, S.; Reisinger, F.; Zoller, J.; Yuan, D.; Heikenwalder, M. Lymphotoxin, NF-κB, and cancer:
The dark side of cytokines. Dig. Dis. 2012, 30, 453–468. [CrossRef] [PubMed]
12. McCarthy, D.D.; Summers-Deluca, L.; Vu, F.; Chiu, S.; Gao, Y.; Gommerman, J.L. The lymphotoxin pathway:
Beyond lymph node development. Immunol. Res. 2006, 35, 41–54. [CrossRef]
13. Upadhyay, V.; Fu, Y.X. Lymphotoxin signalling in immune homeostasis and the control of microorganisms.
Nat. Rev. Immunol. 2013, 13, 270–279. [CrossRef] [PubMed]
14. Kruglov, A.A.; Grivennikov, S.I.; Kuprash, D.V.; Winsauer, C.; Prepens, S.; Seleznik, G.M.; Eberl, G.;
Littman, D.R.; Heikenwalder, M.; Tumanov, A.V. Nonredundant function of soluble LTα3 produced by
innate lymphoid cells in intestinal homeostasis. Science 2013, 342, 1243–1246. [CrossRef] [PubMed]
15. Borgo, F.; Riva, A.; Benetti, A.; Casiraghi, M.C.; Bertelli, S.; Garbossa, S.; Anselmetti, S.; Scarone, S.;
Pontiroli, A.E.; Morace, G.; et al. Microbiota in anorexia nervosa: The triangle between bacterial species,
metabolites and psychological tests. PLoS ONE 2017, 12, e0179739. [CrossRef] [PubMed]
16. Herpertz-Dahlmann, B.; Seitz, J.; Baines, J. Food matters: How the microbiome and gut-brain interaction
might impact the development and course of anorexia nervosa. Eur. Child Adolesc. Psychiatry 2017,
26, 1031–1041. [CrossRef] [PubMed]
21
Nutrients 2018, 10, 1573
17. Lam, Y.Y.; Maguire, S.; Palacios, T.; Caterson, I.D. Are the gut bacteria telling us to eat or not to eat? Reviewing
the role of gut microbiota in the etiology, disease progression and treatment of eating disorders. Nutrients
2017, 9, E602. [CrossRef] [PubMed]
18. Jha, M.K.; Minhajuddin, A.; Gadad, B.S.; Greer, T.L.; Mayes, T.L.; Trivedi, M.H. Interleukin 17 selectively
predicts better outcomes with bupropion-SSRI combination: Novel T cell biomarker for antidepressant
medication selection. Brain Behav. Immun. 2017, 66, 103–110. [CrossRef] [PubMed]
19. Borovcanin, M.; Jovanovic, I.; Radosavljevic, G.; Djukic Dejanovic, S.; Bankovic, D.; Arsenijevic, N.;
Lukic, M.L. Elevated serum level of type-2 cytokine and low IL-17 in first episode psychosis and
schizophrenia in relapse. J. Psychiatr. Res. 2012, 46, 1421–1426. [CrossRef] [PubMed]
20. Himmerich, H.; Schonherr, J.; Fulda, S.; Sheldrick, A.J.; Bauer, K.; Sack, U. Impact of antipsychotics on
cytokine production in-vitro. J. Psychiatr. Res. 2011, 45, 1358–1365. [CrossRef] [PubMed]
21. Turner, M.D.; Nedjai, B.; Hurst, T.; Pennington, D.J. Cytokines and chemokines: At the crossroads of cell
signalling and inflammatory disease. Biochim. Biophys. Acta 2014, 1843, 2563–2582. [CrossRef] [PubMed]
22. Borish, L.C.; Steinke, J.W. 2. Cytokines and chemokines. J. Allergy Clin. Immunol. 2003, 111, S4604–S4675.
[CrossRef]
23. Pisetsky, D.S.; Trace, S.E.; Brownley, K.A.; Hamer, R.M.; Zucker, N.L.; Roux-Lombard, P.; Dayer, J.M.;
Bulik, C.M. The expression of cytokines and chemokines in the blood of patients with severe weight loss
from anorexia nervosa: An exploratory study. Cytokine 2014, 69, 110–115. [CrossRef] [PubMed]
24. Zhang, S.; Tang, H.; Gong, C.; Liu, J.; Chen, J. Assessment of serum CX3CL1/fractalkine level in Han Chinese
girls with anorexia nervosa and its correlation with nutritional status: A preliminary cross-sectional study.
J. Investig. Med. 2017, 65, 333–337. [CrossRef] [PubMed]
25. Murphy, K.; Travers, P.; Walport, M. Janeway’s Immunobiology, 7th ed.; Garland Science, Taylor & Francis
Group: Abingdon, UK, 2009.
26. Víctor, V.M.; Rovira-Llopis, S.; Saiz-Alarcon, V.; Sanguesa, M.C.; Rojo-Bofill, L.; Banuls, C.; de Pablo, C.;
Alvarez, A.; Rojo, L.; Rocha, M.; et al. Involvement of leucocyte/endothelial cell interactions in anorexia
nervosa. Eur. J. Clin. Investig. 2015, 45, 670–678. [CrossRef] [PubMed]
27. Jeon, S.W.; Kim, Y.K. Neuroinflammation and cytokine abnormality in major depression: Cause or
consequence in that illness? World J. Psychiatry 2016, 6, 283–293. [CrossRef] [PubMed]
28. Capuron, L.; Miller, A.H. Immune system to brain signaling: Neuropsychopharmacological implications.
Pharmacol. Ther. 2011, 130, 226–238. [CrossRef] [PubMed]
29. Stuart, M.J.; Singhal, G.; Baune, B.T. Systematic review of the neurobiological relevance of chemokines to
psychiatric disorders. Front. Cell. Neurosci. 2015, 9, 357. [CrossRef] [PubMed]
30. Kowalska, I.; Karczewska-Kupczewska, M.; Straczkowski, M. Adipocytokines, gut hormones and growth
factors in anorexia nervosa. Clin. Chim. Acta 2011, 412, 1702–1711. [CrossRef] [PubMed]
31. Himmerich, H.; Treasure, J. Psychopharmacological advances in eating disorders. Expert Rev. Clin. Pharmacol.
2018, 11, 95–108. [CrossRef] [PubMed]
32. Tong, L.; Prieto, G.A.; Kramár, E.A.; Smith, E.D.; Cribbs, D.H.; Lynch, G.; Cotman, C.W. BDNF-dependent
synaptic plasticity is suppressed by IL-1β via p38 MAPK. J. Neurosci. 2012, 32, 17714–17724. [CrossRef]
[PubMed]
33. Licht, T.; Keshet, E. Delineating multiple functions of VEGF-A. in the adult brain. Cell. Mol. Life Sci. 2013,
70, 1727–1737. [CrossRef] [PubMed]
34. Calabrese, F.; Rossetti, A.C.; Racagni, G.; Gass, P.; Riva, M.A.; Molteni, R. Brain-derived neurotrophic factor:
A bridge between inflammation and neuroplasticity. Front. Cell. Neurosci. 2014, 8, 430. [CrossRef] [PubMed]
35. Donzis, E.J.; Tronson, N.C. Modulation of learning and memory by cytokines: Signaling mechanisms and
long term consequences. Neurobiol. Learn. Mem. 2014, 115, 68–77. [CrossRef] [PubMed]
36. Dugué, B.; Leppanen, E.; Grasbeck, R. Preanalytical factors and the measurement of cytokines in human
subjects. Int. J. Clin. Lab. Res. 1996, 26, 99–105. [CrossRef] [PubMed]
37. Anzengruber, D.; Klump, K.L.; Thornton, L.; Brandt, H.; Crawford, S.; Fichter, M.M.; Halmi, K.A.; Johnson, C.;
Kaplan, A.S.; LaVia, M.; et al. Smoking in eating disorders. Eat. Behav. 2006, 7, 291–299. [CrossRef] [PubMed]
38. Dowlati, Y.; Herrmann, N.; Swardfager, W.; Liu, H.; Sham, L.; Reim, E.K.; Lanctot, K.L. A meta-analysis of
cytokines in major depression. Biol. Psychiatry 2010, 67, 446–457. [CrossRef] [PubMed]
22
Nutrients 2018, 10, 1573
39. Ulfvebrand, S.; Birgegard, A.; Norring, C.; Hogdahl, L.; von Hausswolff-Juhlin, Y. Psychiatric comorbidity
in women and men with eating disorders results from a large clinical database. Psychiatry Res. 2015,
230, 294–299. [CrossRef] [PubMed]
40. Keyes, A.; Woerwag-Mehta, S.; Bartholdy, S.; Koskina, A.; Middleton, B.; Connan, F.; Webster, P.; Schmidt, U.;
Campbell, I.C. Physical activity and the drive to exercise in anorexia nervosa. Int. J. Eat. Disord. 2015,
48, 46–54. [CrossRef] [PubMed]
41. Schmidt, U.; Sharpe, H.; Bartholdy, S.; Bonin, E.M.; Davies, H.; Easter, A.; Goddard, E.; Hibbs, R.; House, J.;
Keyes, A.; et al. Treatment of anorexia nervosa: A multimethod investigation translating experimental
neuroscience into clinical practice. In Programme Grants for Applied Research; NIHR Journals Library:
Southampton, UK, 2017.
42. First, M.B.; Gibbon, M.; Spitzer, R.L.; Williams, J.B.W. User’s Guide for the Structured Clinical Interview for
DSM-IV Axis I Disorders—Research Version; Biometrics Research Department, New York State Psychiatric
Institute: New York, NY, USA, 1996.
43. Fairburn, C. Appendix: Eating Disorder Examination Questionnaire (EDE-Q. Version 6.0) In Cognitive Behaviour
Therapy and Eating Disorders; Guilford Press: New York, NY, USA, 2008.
44. Lovibond, S.; Lovibond, P. Manual for the Depression Anxiety Stress Scales, 2nd ed.; Psychology Foundation:
Sydney, Australia, 1995.
45. StataCorp. Stata Statistical Software: Release 15; StataCorp LLC: College Station, TX, USA, 2017.
46. Luce, K.H.; Crowther, J.H.; Pole, M. Eating Disorder Examination Questionnaire (EDE-Q): Norms for
undergraduate women. Int. J. Eat. Disord. 2008, 41, 273–276. [CrossRef] [PubMed]
47. Mond, J.M.; Hay, P.J.; Rodgers, B.; Owen, C. Eating Disorder Examination Questionnaire (EDE-Q): Norms
for young adult women. Behav. Res. Ther. 2006, 44, 53–62. [CrossRef] [PubMed]
48. Marcos, A. The immune system in eating disorders: An overview. Nutrition 1997, 13, 853–862. [CrossRef]
49. Slotwinska, S.M.; Slotwinski, R. Immune disorders in anorexia. Cent. Eur. J. Immunol. 2017, 42, 294–300.
[CrossRef] [PubMed]
50. Omodei, D.; Pucino, V.; Labruna, G.; Procaccini, C.; Galgani, M.; Perna, F.; Pirozzi, D.; De Caprio, C.;
Marone, G.; Fontana, L.; et al. Immune-metabolic profiling of anorexic patients reveals an anti-oxidant and
anti-inflammatory phenotype. Metabolism 2015, 64, 396–405. [CrossRef] [PubMed]
51. Mörkl, S.; Lackner, S.; Meinitzer, A.; Mangge, H.; Lehofer, M.; Halwachs, B.; Gorkiewicz, G.; Kashofer, K.;
Painold, A.; Holl, A.K.; et al. Gut microbiota, dietary intakes and intestinal permeability reflected by serum
zonulin in women. Eur. J. Nutr. 2018. [CrossRef] [PubMed]
52. Alam, R.; Abdolmaleky, H.M.; Zhou, J.R. Microbiome, inflammation, epigenetic alterations, and mental
diseases. Am. J. Med. Genet. B Neuropsychiatr. Genet. 2017, 174, 651–660. [CrossRef] [PubMed]
53. Żak-Gołab,
˛ A.; Kocelak, P.; Aptekorz, M.; Zientara, M.; Juszczyk, L.; Martirosian, G.; Chudek, J.;
Olszanecka-Glinianowicz, M. Gut microbiota, microinflammation, metabolic profile, and zonulin
concentration in obese and normal weight subjects. Int. J. Endocrinol. 2013, 2013, 674106. [CrossRef]
[PubMed]
54. Pan, W.; Wu, X.; He, Y.; Hsuchou, H.; Huang, E.Y.; Mishra, P.K.; Kastin, A.J. Brain interleukin-15 in
neuroinflammation and behavior. Neurosci. Biobehav. Rev. 2013, 37, 184–192. [CrossRef] [PubMed]
55. Wu, X.; Hsuchou, H.; Kastin, A.J.; He, Y.; Khan, R.S.; Stone, K.P.; Cash, M.S.; Pan, W. Interleukin-15 affects
serotonin system and exerts antidepressive effects through IL15Rα receptor. Psychoneuroendocrinology 2011,
36, 266–278. [CrossRef] [PubMed]
56. Gauthier, C.; Hassler, C.; Mattar, L.; Launay, J.M.; Callebert, J.; Steiger, H.; Melchior, J.C.; Falissard, B.;
Berthoz, S.; Mourier-Soleillant, V.; et al. Symptoms of depression and anxiety in anorexia nervosa: Links
with plasma tryptophan and serotonin metabolism. Psychoneuroendocrinology 2014, 39, 170–178. [CrossRef]
[PubMed]
57. Kaye, W.H.; Frank, G.K.; Bailer, U.F.; Henry, S.E.; Meltzer, C.C.; Price, J.C.; Mathis, C.A.; Wagner, A. Serotonin
alterations in anorexia and bulimia nervosa: New insights from imaging studies. Physiol. Behav. 2005,
85, 73–81. [CrossRef] [PubMed]
58. de Witte, L.; Tomasik, J.; Schwarz, E.; Guest, P.C.; Rahmoune, H.; Kahn, R.S.; Bahn, S. Cytokine alterations in
first-episode schizophrenia patients before and after antipsychotic treatment. Schizophr. Res. 2014, 154, 23–29.
[CrossRef] [PubMed]
23
Nutrients 2018, 10, 1573
59. Bulik-Sullivan, B.; Finucane, H.K.; Anttila, V.; Gusev, A.; Day, F.R.; Loh, P.R.; Duncan, L.; Perry, J.R.;
Patterson, N.; Robinson, E.B.; et al. An atlas of genetic correlations across human diseases and traits.
Nat. Genet. 2015, 47, 1236–1241. [CrossRef] [PubMed]
60. Pedersen, B.K. Muscles and their myokines. J. Exp. Biol. 2011, 214, 337–346. [CrossRef] [PubMed]
61. Kadasah, S.; Arfin, M.; Rizvi, S.; Al-Asmari, M.; Al-Asmari, A. Tumor necrosis factor-α and -β genetic
polymorphisms as a risk factor in Saudi patients with schizophrenia. Neuropsychiatr. Dis. Treat. 2017,
13, 1081–1088. [CrossRef] [PubMed]
62. Ferrara, N. Vascular endothelial growth factor: Basic science and clinical progress. Endocr. Rev. 2004,
25, 581–611. [CrossRef] [PubMed]
63. Sharma, A.N.; da Costa e Silva, B.F.; Soares, J.C.; Carvalho, A.F.; Quevedo, J. Role of trophic factors GDNF,
IGF-1 and VEGF in major depressive disorder: A comprehensive review of human studies. J. Affect. Disord.
2016, 197, 9–20. [CrossRef] [PubMed]
64. Frydecka, D.; Krzystek-Korpacka, M.; Lubeiro, A.; Stramecki, F.; Stanczykiewicz, B.; Beszlej, J.A.;
Piotrowski, P.; Kotowicz, K.; Szewczuk-Boguslawska, M.; Pawlak-Adamska, E.; et al. Profiling inflammatory
signatures of schizophrenia: A cross-sectional and meta-analysis study. Brain Behav. Immun. 2018, 71, 28–36.
[CrossRef] [PubMed]
65. Kluge, M.; Schuld, A.; Schacht, A.; Himmerich, H.; Dalal, M.A.; Wehmeier, P.M.; Hinze-Selch, D.; Kraus, T.;
Dittmann, R.W.; Pollmacher, T. Effects of clozapine and olanzapine on cytokine systems are closely linked to
weight gain and drug-induced fever. Psychoneuroendocrinology 2009, 34, 118–128. [CrossRef] [PubMed]
66. Himmerich, H.; Au, K.; Dornik, J.; Bentley, J.; Schmidt, U.; Treasure, J. Olanzapine treatment for patients
with anorexia nervosa. Can. J. Psychiatry 2017, 62, 506–507. [CrossRef] [PubMed]
67. Davey, K.J.; Cotter, P.D.; O’Sullivan, O.; Crispie, F.; Dinan, T.G.; Cryan, J.F.; O’Mahony, S.M. Antipsychotics
and the gut microbiome: Olanzapine-induced metabolic dysfunction is attenuated by antibiotic
administration in the rat. Transl. Psychiatry 2013, 3, e309. [CrossRef] [PubMed]
68. Davey, K.J.; O’Mahony, S.M.; Schellekens, H.; O’Sullivan, O.; Bienenstock, J.; Cotter, P.D.; Dinan, T.G.;
Cryan, J.F. Gender-dependent consequences of chronic olanzapine in the rat: Effects on body weight,
inflammatory, metabolic and microbiota parameters. Psychopharmacology 2012, 221, 155–169. [CrossRef]
[PubMed]
69. Flowers, S.A.; Evans, S.J.; Ward, K.M.; McInnis, M.G.; Ellingrod, V.L. Interaction between atypical
antipsychotics and the gut microbiome in a bipolar disease cohort. Pharmacotherapy 2017, 37, 261–267.
[CrossRef] [PubMed]
70. Morgan, A.P.; Crowley, J.J.; Nonneman, R.J.; Quackenbush, C.R.; Miller, C.N.; Ryan, A.K.; Bogue, M.A.;
Paredes, S.H.; Yourstone, S.; Carroll, I.M.; et al. The antipsychotic Olanzapine interacts with the gut
microbiome to cause weight gain in mouse. PLoS ONE 2014, 9, e115225. [CrossRef] [PubMed]
71. Wong, S.; Pinkney, J. Role of cytokines in regulating feeding behaviour. Curr. Drug Targets 2004, 5, 251–263.
[CrossRef] [PubMed]
72. Ohsumi, J.; Sakakibara, S.; Yamaguchi, J.; Miyadai, K.; Yoshioka, S.; Fujiwara, T.; Horikoshi, H.; Serizawa, N.
Troglitazone prevents the inhibitory effects of inflammatory cytokines on insulin-induced adipocyte
differentiation in 3T3-L1 cells. Endocrinology 1994, 135, 2279–2282. [CrossRef] [PubMed]
73. Wittchen, E.S. Endothelial signaling in paracellular and transcellular leukocyte transmigration. Front. Biosci.
2009, 14, 2522–2545. [CrossRef]
74. Allavena, R.; Noy, S.; Andrews, M.; Pullen, N. CNS elevation of vascular and not mucosal addressin cell
adhesion molecules in patients with multiple sclerosis. Am. J. Pathol. 2010, 176, 556–562. [CrossRef]
[PubMed]
75. Bauer, M.E.; Teixeira, A.L. Inflammation in psychiatric disorders: What comes first? Ann. N. Y. Acad. Sci. 2018.
[CrossRef] [PubMed]
76. Brandys, M.K.; Kas, M.J.H.; van Elburg, A.A.; Campbell, I.C.; Adan, R.A.H. A meta-analysis of circulating
BDNF concentrations in anorexia nervosa. World J. Biol. Psychiatry 2011, 12, 444–454. [CrossRef] [PubMed]
77. Burkert, N.T.; Koschutnig, K.; Ebner, F.; Freidl, W. Structural hippocampal alterations, perceived stress, and
coping deficiencies in patients with anorexia nervosa. Int. J. Eat. Disord. 2015, 48, 670–676. [CrossRef]
[PubMed]
24
Nutrients 2018, 10, 1573
78. Connan, F.; Murphy, F.; Connor, S.E.; Rich, P.; Murphy, T.; Bara-Carill, N.; Landau, S.; Krljes, S.; Ng, V.;
Williams, S.; et al. Hippocampal volume and cognitive function in anorexia nervosa. Psychiatry Res. 2006,
146, 117–125. [CrossRef] [PubMed]
79. Audet, M.-C.; Anisman, H. Interplay between pro-inflammatory cytokines and growth factors in depressive
illnesses. Front. Cell. Neurosci. 2013, 7, 68. [CrossRef] [PubMed]
80. Canavan, B.; Salem, R.O.; Schurgin, S.; Koutkia, P.; Lipinska, I.; Laposata, M.; Grinspoon, S. Effects of
physiological leptin administration on markers of inflammation, platelet activation, and platelet aggregation
during caloric deprivation. J. Clin. Endocrinol. Metab. 2005, 90, 5779–5785. [CrossRef] [PubMed]
81. Raschke, S.; Eckel, J. Adipo-myokines: Two sides of the same coin–mediators of inflammation and mediators
of exercise. Mediat. Inflamm. 2013, 2013, 320724. [CrossRef] [PubMed]
82. Kyle, U.G.; Bosaeus, I.; De Lorenzo, A.D.; Deurenberg, P.; Elia, M.; Manuel Gomez, J.; Lilienthal Heitmann, B.;
Kent-Smith, L.; Melchior, J.C.; Pirlich, M.; et al. Bioelectrical impedance analysis-part II: Utilization in clinical
practice. Clin. Nutr. 2004, 23, 1430–1453. [CrossRef] [PubMed]
83. Ackland, T.R.; Lohman, T.G.; Sundgot-Borgen, J.; Maughan, R.J.; Meyer, N.L.; Stewart, A.D.; Muller, W.
Current status of body composition assessment in sport: Review and position statement on behalf of the ad
hoc research working group on body composition health and performance, under the auspices of the I.O.C.
Medical Commission. Sports Med. 2012, 42, 227–249. [CrossRef] [PubMed]
84. Labrecque, N.; Cermakian, N. Circadian clocks in the immune system. J. Biol. Rhythms 2015, 30, 277–290.
[CrossRef] [PubMed]
85. Munzer, A.; Sack, U.; Mergl, R.; Schonherr, J.; Petersein, C.; Bartsch, S.; Kirkby, K.C.; Bauer, K.; Himmerich, H.
Impact of antidepressants on cytokine production of depressed patients in vitro. Toxins 2013, 5, 2227–2240.
[CrossRef] [PubMed]
© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
25
nutrients
Article
TGF-β2, EGF, and FGF21 Growth Factors Present in
Breast Milk Promote Mesenteric Lymph Node
Lymphocytes Maturation in Suckling Rats
Paulina Torres-Castro 1,2 , Mar Abril-Gil 1,2 , María J. Rodríguez-Lagunas 1,2 , Margarida Castell 1,2 ,
Francisco J. Pérez-Cano 1,2, * and Àngels Franch 1,2
1 Physiology Section, Department of Biochemistry and Physiology, Faculty of Pharmacy and Food Science,
University of Barcelona, 08028 Barcelona, Spain; mtorreca29@alumnes.ub.edu (P.T.-C.);
mariadelmar.abril@ub.edu (M.A.-G.); mjrodriguez@ub.edu (M.J.R.-L.); margaridacastell@ub.edu (M.C.);
angelsfranch@ub.edu (À.F.)
2 Nutrition and Food Safety Research Institute (INSA·UB), 08921 Santa Coloma de Gramenet, Spain
* Correspondence: franciscoperez@ub.edu; Tel.: +34-934-024-505
Abstract: Breast milk, due to its large number of nutrients and bioactive factors, contributes to
optimal development and immune maturation in early life. In this study, we aimed to assess the
influence of some growth factors present in breast milk, such as transforming growth factor-β2
(TGF-β2), epidermal growth factor (EGF), and fibroblast growth factor 21 (FGF21), on the immune
response development. Newborn Wistar rats were supplemented daily with TGF-β2, EGF, or FGF21,
throughout the suckling period. At day 14 and 21 of life, lymphocytes from mesenteric lymph nodes
(MLNs) were isolated, immunophenotyped, and cultured to evaluate their ability to proliferate
and release cytokines. The main results demonstrated that supplementation with TGF-β2, EGF,
or FGF21 modified the lymphocyte composition in MLNs. At day 14, all supplementations were able
to induce a lower percentage of natural killer (NK) cells with the immature phenotype (CD8+ ), and
they reduced the CD8αα/CD8αβ ratio at day 21. Moreover, the cytokine pattern was modified by
the three treatments, with a down regulation of interleukin (IL)-13 secretion. These results showed
the contribution of these growth factors in the lymphocytes MLNs immune maturation during the
neonatal period.
1. Introduction
At time of birth, the intestine is immature, not only anatomically, but also metabolically and
immunologically [1–3]. Intestinal development is a key process in early life because it includes
important functions related to growth and survival. It is important to develop mechanisms to digest
and absorb nutrients in an efficient way [4]. Moreover, the intestine begins hosting the gut microbiota
and establishing appropriate host immune responses against pathogens [2]. The intestinal maturation
process can be conditioned through synergy of several factors, such as genetics, microbial colonization,
and nutrition [1–3].
Breast milk is the gold standard to feed the newborn because it includes a rich number of
components, which are essential for optimal growth and development [5]. It also contains a high
number of bioactive factors, which participate in the immune maturation process of infants [2,6].
Among these components, immunoglobulins (Igs), cytokines, and growth factors (GFs) have an
important role [7–9]. In this sense, the effects of several GFs present in maternal milk that promote
intestinal maturation have been described, although their impact on immune development is still
unclear [9].
On the one hand, the transforming growth factor-β (TGF-β) family members have multifunctional
actions involved in maintaining intestinal homeostasis, regulating inflammation, and allergy
development [10–12]. TGF-βs act on different types of leukocytes, to control the initiation and
resolution of immune responses through the recruitment, activation, and survival of cells [12,13].
TGF-β2 is the predominant isoform present in human and rat breast milk; it reaches the neonatal
intestine where, at birth, endogenous production is low and increases towards weaning [3,10,12].
In infants, breast milk TGF-βs play an important role in developing immune response and promoting
oral tolerance development [12].
On the other hand, one of the most abundant GFs in breast milk is the epidermal growth factor
(EGF), the concentration of which is 500 times more than other GFs present in breast milk [14].
Its concentration is very high in colostrum and it decreases significantly during suckling, both in
human and rodent milk, suggesting that EGF plays a role in the promotion of early neonatal intestinal
growth [2,15]. In fact, EGF is a key intestinal regulator in protecting intestinal barrier integrity, essential
for the absorption of nutrients and the exclusion of pathogens, in both humans and animals [14,16,17].
EGF is a polyfacetic molecule, which acts by regulating different processes, such as cell growth,
survival, migration, apoptosis, proliferation, and differentiation. In early life, milk EGF seems to be
one of the crucial components involved in the prevention of necrotizing enterocolitis (NEC) [1,14].
EGF and TGF-β together with the immunosuppressive interleukin (IL)-10, also present in breast
milk, are involved in the functional development of the gastrointestinal mucosa, tolerance acquisition,
and inflammation downregulation in damaged intestinal cells [6,7].
In recent years, new components present in breast milk have been discovered. This is the case
of the fibroblast growth factor 21 (FGF21), which belongs to the hormone-like subgroup within the
FGF superfamily, and has been found in rodent and human milk [16]. The FGF21 present in milk,
seems to be involved in local actions in the neonatal intestine. It is a highly active pleiotropic factor,
involved in multiple aspects of metabolism through a variety of mechanisms, where it regulates both
the expression and activity of digestive enzymes, and the synthesis and release of several intestinal
hormone factors [16,18]. All these effects make FGF21 a good candidate to be studied, due to its
possible contribution to neonatal intestinal function.
Overall, because TGF-β2, EGF, and FGF21 are biologically active factors found in breast milk
and are suggested to be involved in neonatal intestine maturation; our hypothesis is that they could
also have an important role in the immune development process in early life. Therefore, the main
objective of the current study was to ascertain whether the effect of a daily supplementation with
these compounds could promote immune response development, during the suckling period in rats.
The effects of the supplementations on mesenteric lymph nodes (MLNs), an inductor site of the gut
associated lymphoid tissue (GALT), were studied specifically during and at the end of the suckling
period, in terms of lymphocyte phenotypic composition and on the immune functionality, such as their
lymphoproliferative and cytokine production abilities.
2.1. Animals
Pregnant Wistar rats (G15) were obtained from Janvier Labs (Le Genest-Saint-Isle, France), and
were individually housed in cages under controlled conditions of temperature and humidity in a
12:12 h light:dark cycle with access to food and water ad libitum. The pregnant rats were monitored
daily and allowed to deliver naturally. The day after birth was reported as day 1. The studies were
performed in accordance with institutional guidelines, for the care and use of laboratory animals, and
were approved by the Ethical Committee for Animal Experimentation (CEEA) at the University of
Barcelona (UB) and Catalonia Government (CEEA-UB Ref. 220/15, UB/DAAM 8521).
27
Nutrients 2018, 10, 1171
28
Nutrients 2018, 10, 1171
MO, USA), 1% L-glutamine (Sigma-Aldrich, St. Louis, MO, USA), 1% penicillin streptomicin (PenStrep;
Sigma-Aldrich, St. Louis, MO, USA), and 0.05 mM 2-β-mercaptoethanol (Merck, Darmstadt, Germany).
MLNs lymphocytes were obtained in sterile conditions by passing the tissue through a cell strainer
(40 μm, BD Biosciences, San Diego, CA, USA). The cell suspensions were centrifuged (538 g, 10 min,
4 ◦ C), and the pellet was resuspended with complete RPMI medium. The cell counts and viabilities
were determined using an automated cell counter, after staining the cells with trypan blue (CountessTM ,
Invitrogen, Madrid, Spain), following usual laboratory procedures, as described in Reference [22].
The lymphocytes were immediately used to characterize their phenotype, and to determine their
ability to proliferate and secrete cytokines.
29
Nutrients 2018, 10, 1171
in combination with the TCRγδ+ cells (obtained from Panel 2) constituted the total of T cells.
B cells (CD45RA+ ) were identified with Panel 4. The mAbs used in this study are detailed in the
Supplementary Materials (Table S1).
Briefly, cells were incubated with a mixture of 10 μL of saturating concentrations of each mouse
anti-rat mAbs in PBS pH 7.2, containing 2% FBS and 0.1% sodium azide (Merck, Darmstadt, Germany),
at 4 ◦ C in darkness for 20 min. For T reg evaluation, an intracellular staining was performed. For that,
cells previously labeled with anti-CD4-PE and anti-CD25-FITC mAbs, were fixed/permeabilized
using a specific buffer kit (eBioscience, San Diego, CA, USA). Then, intracellular staining with
anti-Foxp3-APC mAb was carried out, under the same conditions as extracellular staining. After
washing, all stained cells were fixed with 0.5% paraformaldehyde (Panreac, Barcelona, Spain) and
stored at 4 ◦ C in darkness, until analysis by flow cytometry. For each sample, a positive control
staining using each isotype matched mAbs and a negative control without staining were included.
Analyses were performed in a Gallios™ Cytometer (Beckman Coulter, Miami, FL, USA) in the
CCiT-UB. Data were assessed by FlowJo® version 10 software (Tree Star Inc., Ashland, Covington,
KY, USA). Results were expressed as percentages of positive cells in the lymphocyte population,
selected according to their forward- and side-scatter characteristics (FSC/SSC) using previous studies
protocols [19], or in a selected population. The gating strategy was specific for each panel used, but
overall, markers (e.g., CD8α in Panel 1, αE integrin, and CD62L in Panel 3) were evaluated in specific
gate subsets (e.g., NK, NKT, or T cells in Panel 1, and CD8 or CD4 cells in Panel 3).
3. Results
30
Nutrients 2018, 10, 1171
All groups increased BMI, SI relative weight, and SI relative length with age (p < 0.05, day 21 vs.
day 14). Regarding the Lee index, only the animals from the REF group and those supplemented with
TGF-β2, showed significant differences associated with age (p < 0.05), but the supplementation with
EGF and FGF21 did not display such a significant age difference.
Although at 14 days there were no differences among the supplemented groups, some differences
could be seen at 21 days (Table 2). At the end of the suckling period, only the EGF supplementation
showed a significant decrease in the percentage of TCRγδ+ cells, compared to the REF group (p < 0.05).
Moreover, the REF and groups supplemented with TGF-β2 and EGF decreased the NK cell proportion,
in comparison to the same group at 14 days (p < 0.05). The animals from the REF and FGF21 groups
increased the percentage of NKT cells with age, and only those from the EGF group were able to
decrease the proportion of TCRγδ+ and B cell percentages at day 21 (p < 0.05 vs. day 14).
The percentage of CD8+ cells and the CD8αα/CD8αβ ratio from MLNs lymphocytes, were
determined as indicators of immune maturation (Figure 1). The proportion of CD8+ cells did not show
significant differences, either associated with age (day 14 vs. day 21) or to any of the supplementations,
compared to the REF group (Figure 1a). However, regarding the results from CD8αα/CD8αβ ratio, the
EGF supplementation induced a significant increase at 14 days (p < 0.05 vs. REF, Figure 1b). At day 21,
in all supplemented groups, the CD8αα/CD8αβ ratio had decreased when they were compared to
their respective group at day 14, but no differences among the interventions and the REF groups were
found (Figure 1b).
31
Nutrients 2018, 10, 1171
&'αα/&'αβ
Ψ Ψ
Ψ
Figure 1. Proportion of CD8+ cells and ratio CD8αα/CD8αβ in mesenteric lymph nodes at 14 and
21 days of life. (a) % CD8+ cells; (b) The ratio CD8αα/CD8αβ was calculated as the quotient of
the percentages of CD8αα cells and CD8αβ cells in CD8+ cells (%). Results are expressed as mean
± standard error of the mean (S.E.M) (n = 9). Statistical differences: * p < 0.05 vs. reference group
(REF group) at same age; Ψ p < 0.05 vs. same group at day 14. TGF-β2: transforming growth factor-β2.
EGF: epidermal growth factor. FGF21: fibroblast growth factor 21.
Further analysis of CD8+ and CD8− subsets revealed some changes associated with the
supplementations (Figure 2).
Regarding the TCRαβ+ subsets (CD8+ and CD8− , respectively, Figure 2a,b), no changes due
to supplementation were found either at 14 or at 21 days, with respect to the REF group. Only, the
supplementation with FGF21 decreased the proportion of TCRαβ+ CD8− (~8%), when values from
day 21 were compared to day 14 (Figure 2a, p < 0.05). The TCRγδ+ CD8− cell percentage (Figure 2c),
increased due to supplementation with EGF and FGF21 at 14 days, with only the latter achieving
statistical significance, but these changes were not observed at the end of the suckling period (day 21).
However, both interventions showed a 50% decrease in the TCRγδ+ CD8− cell proportion associated
with age (p < 0.05, day 21 vs. day 14). Moreover, the TCRγδ+ CD8+ cell proportion in the EGF group
(Figure 2d), was lower than the REF group at day 21 (p < 0.05). In relation to the NKT population (CD8-
and CD8+ subsets), only the NKT CD8− subset showed changes associated with age for all groups
(Figure 2e), but not due to the supplementations.
32
Nutrients 2018, 10, 1171
7&5αβ &'
7&5αβ + &' ψ
(c) (d)
7&5γδ&'
7&5γδ&'
ψ
ψ
(e) (f)
ψ
ψ
1.7&'
ψ ψ
1.7&'
(g) (h)
1.&'
1.&'
ψ
ψ ψ
ψ
Figure 2. Percentages of CD8+ and CD8− lymphocyte subsets in mesenteric lymph nodes at 14 and
21 days of life. (a) TCRαβ+ CD8− ; (b) TCRαβ+ CD8+ ; (c) TCRγδ+ CD8− ; (d) TCRγδ+ CD8+ ; (e) NKT
CD8− ; (f) NKT CD8+ ; (g) NK CD8- ; and (h) NK CD8+ . Results are expressed as mean ± S.E.M (n = 9).
Statistical difference: * p < 0.05 vs. REF group at same age; Ψ p < 0.05 vs. same group at day 14.
However, NKT CD8+ cell proportions were not affected (Figure 2f). Results from the NK cell
population showed that in supplemented groups, there was a decrease in NK CD8- cell proportions
related to age (Figure 2g, p < 0.05, day 21 vs. day 14), and that only supplementation with TGF-β2
induced a significant decrease compared with the REF group at 21 days (Figure 2g, p < 0.05 vs. REF).
The phenotype of the intestinal NK cells based on CD8 expression, could be considered as immature
(NK CD8+ ) or more mature (NK CD8− ). The three supplementations were able to decrease the
proportion of NK cells expressing CD8 at 14 days (Figure 2h), which reached similar values to those
from reference 21-day-old rats. Thus, only the REF group decreased its percentage of NK CD8+ cells at
21 days, with respect to day 14 (Figure 2h).
33
Nutrients 2018, 10, 1171
The lymphocytes’ commitment to the mucosal compartment were studied by means of the
proportion of cells expressing two adhesion molecules of importance in the intestinal homing; thus, the
total percentage of cells bearing the selectin CD62L and the αE integrin were determined (Figure 3).
Figure 3. Surface expression of αE integrin and CD62L selectin, in mesenteric lymph nodes at 14 and
21 days of life. (a) CD62L/αE integrin surface expression at 14 days, (b) CD62L/αE integrin surface
expression at 21 days, (c) CD62L+ cells (%) and (d) αE+ cells (%). Results are expressed as mean ±
S.E.M (n = 9). Statistical difference: Ψ p > 0.05 vs. same group at day 14.
Figure 3a,b show the molecular pattern of αE/CD62L on day 14 and 21. The CD62L molecule was
expressed on both days in high proportion of cells (~60%). Although the percentage of CD62L+ cells
increased significantly with age in all studied groups (Figure 3a–c, p < 0.05), none of the supplemented
groups induced significant differences, when they were compared to the REF group at the same age.
The αE+ cells were present in a proportion lower than 5% at both studied times (Figure 3b–d) and
were not modified by any GFs administration (Figure 3d).
The percentages of integrin αE+ cells and selectin CD62L+ cells in CD8+ , CD4+ , and B cells were
further studied (Table 3). None of the supplemented groups showed significant differences, compared
to the REF group at the same age. All studied groups increased CD62L+ CD8+ and CD62L+ CD4+
percentages with age (p < 0.05, day 21 vs. day 14). The REF group increased (~15%) αE+ B cells
subset proportion, with respect to the same group at 14 days, whereas all supplemented groups
decreased (~20%) this subset with age (p < 0.05). In addition, the supplementation with EGF and
FGF21, decreased the percentage of αE expression in CD4+ cells with age. Although there was a
tendency to increase with age, the proportion of CD62L+ in B cells in all groups, it was only significant
in the TGF-β2 group (p < 0.05).
At the end of the suckling period, activated CD4 cells (Foxp3− CD25+ CD4+ cells) and Treg cells
(Foxp3+ CD25+ CD4+ cells) were also studied, and no changes were found due to diets (day 21), with
all groups together having a mean percentage of 1.22 ± 0.12 and 3.81 ± 0.11, respectively.
34
Nutrients 2018, 10, 1171
Table 3. Surface expression of the αE integrin and the CD62L selectin in CD4+ , CD8+ , and B cells in
mesenteric lymph nodes.
14 Days
Reference TGF-β2 EGF FGF21
CD8+ 2.75 ± 0.72 2.92 ± 0.49 2.94 ± 0.52 2.88 ± 0.44
%αE CD4+ 2.81 ± 0.39 3.07 ± 0.45 4.46 ± 0.34 3.80 ± 0.26
B cells 15.98 ± 1.61 18.33 ± 1.39 18.92 ± 2.50 20.32 ± 3.12
CD8+ 65.27 ± 1.56 62.38 ± 3.47 59.25 ± 2.52 59.21 ± 3.55
%CD62L CD4+ 59.55 ± 1.76 55.41 ± 2.36 53.61 ± 1.37 53.65 ± 2.40
B cells 48.43 ± 2.77 46.6 ± 2.05 38.77 ± 2.08 37.29 ± 4.78
21 Days
Reference TGF-β2 EGF FGF21
CD8+ 3.84 ± 0.48 2.98 ± 1.04 3.70 ± 0.48 3.97 ± 0.66
%αE CD4+ 2.83 ± 0.75 2.61 ± 0.52 1.94 ± 0.18 Ψ 2.52 ± 0.30 Ψ
B cells 18.43 ± 1.83 Ψ 13.03 ± 1.55 Ψ 14.95 ± 1.69 Ψ 15.75 ± 1.16 Ψ
CD8+ 71.49 ± 1.38 Ψ 72.10 ± 1.93 Ψ 70.78 ± 2.19 Ψ 72.88 ± 2.37 Ψ
%CD62L CD4+ 66.02 ± 1.74 Ψ 68.24 ± 1.64 Ψ 66.66 ± 0.42 Ψ 66.80 ± 1.22 Ψ
B cells 56.79 ± 2.25 60.92 ± 2.17 Ψ 60.82 ± 3.92 57.61 ± 2.37
Results are expressed as mean ± S.E.M (n = 9). Ψ p < 0.05 vs. day 14 at same group.
35
Nutrients 2018, 10, 1171
Table 4. Cytokine production by mesenteric lymph node cells after in vitro stimulation.
14 Days
(pg/mL) Reference TGF-β2 EGF FGF21
IL-2 42.71 ± 5.96 39.24 ± 5.66 50.45 ± 6.66 56.72 ± 11.04
IL-4 89.82 ± 13.00 57.14 ± 7.81 82.65 ± 12.47 74.97 ± 12.18
IL-10 879.82 ± 239.40 688.01 ± 145.14 1450.87 ± 296.29 1326.99 ± 217.06
IL-13 25.38 ± 5.64 17.53 ± 3.03 31.91 ± 4.28 26.18 ± 4.03
IL-17A 71.98 ± 26.92 70.22 ± 20.16 60.12 ± 13.61 50.82 ± 9.53
IFN-γ 10323.26 ± 2391.46 8302.75 ± 1964.09 9329.17 ± 1662.87 7882.51 ± 2810.34
TNF-α 8.76 ± 0.98 9.48 ± 0.68 9.19 ± 1.16 7.40 ± 0.25
IL-10/TNF-α 115.00 ± 33.11 79.71 ± 21.80 170.87 ± 43.10 178.05 ± 29.25
IFN-γ/IL-4 110.35 ± 20.65 195.32 ± 61.39 124.85 ± 37.55 103.80 ± 26.93
21 Days
(pg/mL) Reference TGF-β2 EGF FGF21
IL-2 390.57 ± 141.49 Ψ 383.80 ± 89.21 Ψ 244.79 ± 80.37 Ψ 270.16 ± 114.54
IL-4 70.52 ± 10.51 52.21 ± 7.69 48.52 ± 1.91 Ψ 66.60 ± 8.80
IL-10 705.45 ± 63.95 514.70 ± 55.30 508.26 ± 61.43 Ψ 542.36 ± 78.82 Ψ
IL-13 19.51 ± 2.44 12.92 ± 0.74 * 9.91 ± 1.24 *,Ψ 9.83 ± 1.65 *,Ψ
IL-17A 59.97 ± 7.14 46.91 ± 4.03 52.30 ± 10.7022 41.24 ± 11.77
IFN-γ 4231.26 ± 365.41 Ψ 4488.50 ± 416.14 6900.74 ± 1129.78 4595.11 ± 1058.87
TNF-α 6.63 ± 0.11 Ψ 6.55 ± 0.20 Ψ 10.02 ± 0.40 * 5.91 ± 0.38 Ψ
IL-10/TNF-α 106.47 ± 9.73 80.13 ± 9.55 50.48 ± 5.72 *,Ψ 95.64 ± 17.03 Ψ
IFN-γ/IL-4 64.25 ± 9.61 111.99 ± 25.23 147.25 ± 27.9166 81.50 ± 25.37
Results are expressed as mean ± S.E.M (n = 9). * p < 0.05 vs. REF group at same age; Ψ p < 0.05 vs. day 14 at
same group.
4. Discussion
At birth, the immune system is immature, as evidenced by a poor antibody production and low
proliferative response of immune cells. In addition, there are mucosal immune impairments, such
as low intestinal IgA content, reduced number of B lymphocytes in the intestinal mucosa, and few
intestinal T cells [23]. Immune development is driven, among other factors, by components of breast
milk. It is known that growth factors, such as TGF-β2 and EGF, regulate the immune response in early
life and confer protective effects against gut mucosal inflammation by enhancing oral tolerance [1,2,7].
FGF21, which is also present in maternal milk, is a less-studied component and could be involved
in such effects as well [16]. The current study aimed to evaluate whether supplementation with
TGF-β2, EGF, or FGF21 had a role in immune maturation in early life. A rat pup model was used to
evaluate the effect of a daily supplementation with TGF-β2, EGF, or FGF21 during suckling on the
GALT, particularly in the MLNs, which is an inductor site of intestinal immune response. We assessed
the immune maturation by functions, such as lymphoproliferation, cytokine production ability, and
establishing lymphocyte MLNs composition.
Regarding growth, the body weight of pups was not modified due to supplementation with
any of the GFs studied. These results were in line with other investigations, in which rats receiving
either TGF-β2 [10] or EGF [24,25] during the first two weeks of life did not change their body weight.
Likewise, although a study showed that the body weight of FGF21-knockout mice compared to
wild-type 3-month-old mice was not different [26], its physiological role as a weight regulating factor
was discussed [18]. Overall, it seems that the tested GFs, under the conditions we used, did not have a
key role in the growth of neonates.
Intestinal length and weight are useful tools for evaluating the primary impact of a nutrient
on the maturation of the rat small intestine [27]. The GFs supplementation in the current study did
not affect these variables. However, it has been reported that suckling rats receiving intraperitoneal
administration of EGF (100 μg/kg) for only two days, increased stomach and intestinal weights [24];
and that rat pups receiving oral administration of EGF through formula at concentrations exceeding
36
Nutrients 2018, 10, 1171
the reported concentrations of EGF in rodent milk, enhanced intestinal growth (weight and length) [15].
Thus, in agreement with our results, only high doses of these compounds seem to be able to modify
intestinal growth.
Although no impact on the body and intestinal growth due to any of these GFs was observed,
some effects of TGF-β2, EGF, or FGF21 on immune variables have been shown to be specific; and
therefore, the influence of each GF tested on MLN lymphocyte maturation will be discussed separately.
Transforming growth factor-β (TGF-β) has a wide range of biological activities and among
them, it has an important role in cell proliferation and differentiation. TGF-β acts as a cytokine
having predominantly suppressive effects on the growth of T and B lymphocytes. In this study,
the supplementation of suckling rats with TGF-β2 did not have a significant influence on unspecific
MLNs proliferative cell response. This result contrasted with a report showing that suckling rats
receiving a whey-enriched TGF-β formula, down-regulated the MLNs lymphocyte proliferative
response to specific antigen after being sensitized [28]. On the other hand, we observed that MLNs
cells after TGF-β2 supplementation did not modify the changes in IL-2 and TNF-α secretion associated
with age, but attenuated IL-13 production at day 21 with respect to reference animals. The attenuation
on IL-13 production, a Th2 cytokine linked to allergic processes [29–31], agrees with results showing
that Brown Norway rat pups receiving a formula with TGF-β2 between 4 and 18 days of life shifted the
immune response from a Th2 type towards a Th1 profile [13]. Likewise, it is known that low levels of
IL-13 in colostrum and mature milk are associated with less eczema in early life [32]. Thus, our results
suggested that early TGF-β2 intake could play an important role in the prevention of Th2 mediated
alterations, such as allergy.
In the current study, the TGF-β2 supplementation was already able to modulate MLN lymphocyte
composition in suckling rats. It decreases the proportion of the immature phenotype of the intestinal
NK cells (NK CD8+ ) on day 14 at levels observed later (day 21); which indicated a positive action
on the intestinal immune maturation. This developmental pattern, decreasing CD8 expression in the
NK cell surface with age, has previously described in rat neonatal intraepithelial lymphocytes (IEL)
and MLN cells [19,27]. In line with this, we found that, although no changes in total CD8+ cells were
observed due to TGF-β2 supplementation, a lower CD8αα/CD8αβ ratio at day 21, compared to day
14 appeared. This ratio has been described to be reduced according to age in healthy MLNs cells
from suckling rats [19], thus our results may suggested a promotion of the intestinal immune system
maturation. This contrasts with results in BALB/c mice pups, showing higher CD8+ T cell proportions
when their lactating mothers were treated with mAbs, against TGF-β twice a week from delivery until
weaning [33]. On the other hand, we did not find changes in the MLNs Treg (CD4+ CD25+ Foxp3+ ) cell
proportion. However, some authors have suggested that oral tolerance induced by breastfeeding could
depend on TGF-β signaling from breast milk, which would be able to up-regulate Foxp3+ cells [33].
Regarding the intestinal humoral immune response, it is known that IgA is poorly produced
by the neonate mucosal immune system [1,28], and that TGF-β1 and TGF-β2 from mammalian milk
are able to stimulate the synthesis of mucosal IgA [28,33]. Thus, TGF-β acting in synergy with IL-10,
can promote IgA production and oral tolerance induction [28]. However, our results did not show
significant changes in the intestinal IgA and IgM content of the suckling rats. This lack of effect could
be due to the fact that breast milk contains IgA and IgM, which are transferred to the pups, and then
these milk antibodies in the intestine would mask the pups’ own levels. For this reason, intestinal IgA
and IgM assessment cannot be a good marker of nutritional supplementation at this level.
Epidermal growth factor (EGF) is a peptide that modulates a variety of biological responses, such
as cell proliferation and differentiation [14]. It is known that it has a clear effect on epithelium, where it
accelerates maturation and stimulates cell proliferation [34–36]. However, in the current study, we have
not detected any effect of the EGF supplementation on the proliferative response of MLNs cells in the
suckling pups. There is evidence that EGF prevents and reduces the incidence and severity of NEC
by modulating important transcription factors for cytokine regulation [37], and it is known that the
balance of pro-inflammatory and anti-inflammatory cytokines may play a key role in the development
37
Nutrients 2018, 10, 1171
of NEC [17]. This effect may be attributable to a down-regulation of pro-inflammatory IL-18 and to
an increase of anti-inflammatory IL-10 at the site of injury [17,37]. Here we found that, 21-day EGF
supplementation increased TNF-α and decreased IL-13 production by MLNs cells. The changes in these
two cytokines, contained in maternal milk [32,38], could be associated with positive effects. Indeed,
decreased IL-13 can be useful in the prevention of allergy, as stated for TGF-β2 [29–31]. Moreover,
EGF supplementation kept the levels of TNF-α on day 21 at the same level as that on day 14, which
could play in favor of its own effects because it is known that EGF receptor (EGF-R) is up-regulated by
TNF-α [39]. Regarding MLNs cell composition, EGF supplementation shares some maturate effects
with TGF-β2, such as the induction of lower levels of NK CD8+ cells and CD8αα/CD8αβ ratio.
Finally, focusing on fibroblast growth factor 21 (FGF21), recent studies showed that its function is
not limited to the regulation of metabolism, but that it is also involved in the protection of multiple
physiological processes, such as oxidation, inflammation, atherosclerosis, and aging processes [40].
A recent study demonstrated that, in the spleen of mice with collagen-induced arthritis (CIA), FGF21
reduces the expression of inflammatory cytokines, such as IL-17, TNF-α, IL-1β, IL-6, IL-8, and MMP3,
whereas IL-10 levels were increased, compared to PBS treated CIA mice [41]. In line with this, we have
not found any significant difference with respect to non-supplemented animals on MLNs inflammatory
cytokines, but the FGF21 supplementation decreased IL-13 at 21 days, as we also found with the other
GFs, suggesting, therefore, its role in preventing allergic events.
It is known that FGF21 is present in breast milk and does not contribute to systemic levels in
mouse neonates, but appears to act locally on the mouse neonatal intestine [16]; and it is unknown
whether this factor plays an important role in the maturation of the immune system. We found that
suckling rats receiving FGF21 promoted maturation of the early life intestinal immune system by
accelerating the decrease in the proportion of NK CD8+ cells and the decrease in the CD8αα/CD8αβ
cell ratio in MLNs, as well as TGF-β2 and EGF. This is the first time that the immunomodulatory effect
of FGF21 has been demonstrated in early life.
5. Conclusions
In conclusion, our study demonstrated that supplementation with TGF-β2, EGF, or FGF21 during
the suckling period had an immunoregulatory effect. Although some specific effects appeared, the
three growth factors were able to modulate similar aspects of MLN cells, such as promoting lymphocyte
maturation, as observed by increasing NK cells with a more mature phenotype (CD8− ) and reducing
IL-13 production, which could be useful in avoiding allergic processes. Further studies should be
carried out to establish the effect of the supplementation of TGF-β2, EGF, or FGF21 on the response of
suckling pups in other parts of the intestinal immune system, such as in the intestinal epithelium or
even at the systemic level.
38
Nutrients 2018, 10, 1171
References
1. Turfkruyer, M.; Verhasselt, V. Breast milk and its impact on maturation of the neonatal immune system. Curr.
Opin. Infect. Dis. 2015, 28, 199–206. [CrossRef] [PubMed]
2. García, C.; Duan, R.D.; Brévaut-Malaty, V.; Gire, C.; Millet, V.; Simeoni, U.; Bernard, M.; Armand, M. Bioactive
compounds in human milk and intestinal health and maturity in preterm newborn: An overview. Cell. Mol.
Biol. 2013, 59, 108–131. [CrossRef] [PubMed]
3. Penttila, I.A.; van Spriel, A.B.; Zhang, M.F.; Xian, C.J.; Steeb, C.B.; Cummins, A.G.; Zola, H.; Read, L.C.
Transforming growth factor-β levels in maternal milk and expression in postnatal rat duodenum and ileum.
Pediatr. Res. 1998, 44, 524–531. [CrossRef] [PubMed]
4. Buddington, R.K.; Sangild, P.T. Companion animals symposium: Development of the mammalian
gastrointestinal tract, the resident microbiota, and the role of diet in early life. J. Anim. Sci. 2011, 89,
1506–1519. [CrossRef] [PubMed]
5. Jacobi, S.K.; Odle, J. Nutritional factors influencing intestinal health of the neonate. Adv. Nutr. 2012, 3,
687–696. [CrossRef] [PubMed]
6. Castellote, C.; Casillas, R.; Ramírez-Santana, C.; Pérez-Cano, F.J.; Castell, M.; Moretones, M.G.;
López-Sabater, M.C.; Franch, À. Premature delivery influences the immunological composition of colostrum
and transitional and mature human milk. J. Nutr. 2011, 141, 1181–1187. [CrossRef] [PubMed]
7. Field, C.J. The immunological components of human milk and their effect on immune development in
infants. J. Nutr. 2005, 135, 1–4. [CrossRef] [PubMed]
8. Andreas, N.J.; Kampmann, B.; Mehring Le-Doare, K. Human breast milk: A review on its composition and
bioactivity. Early Hum. Dev. 2015, 91, 629–635. [CrossRef] [PubMed]
9. Ballard, O.; Morrow, A.L. Human milk composition: Nutrients and bioactive factors. Pediatr. Clin. N. Am.
2013, 60, 49–74. [CrossRef] [PubMed]
10. Penttila, I.A.; Flesch, I.E.; McCue, A.L.; Powell, B.C.; Zhou, F.H.; Read, L.C.; Zola, H. Maternal milk regulation
of cell infiltration and interleukin 18 in the intestine of suckling rat pups. Gut 2003, 52, 1579–1586. [CrossRef]
[PubMed]
11. Sanjabi, S.; Oh, S.A.; Li, M.O. Regulation of the immune response by TGF-β: From conception to
autoimmunity and infection. Cold Spring Harb. Perspect. Biol. 2017, 9, a022236. [CrossRef] [PubMed]
12. Penttila, I.A. Milk-derived transforming growth factor-beta and the infant immune response. J. Pediatr. 2010,
156, S21–S25. [CrossRef] [PubMed]
13. Penttila, I. Effects of transforming growth factor-beta and formula feeding on systemic immune responses to
dietary β-lactoglobulin in allergy-prone rats. Pediatr. Res. 2006, 59, 650–655. [CrossRef] [PubMed]
14. Tang, X.; Liu, H.; Yang, S.; Li, Z.; Zhong, J.; Fang, R. Epidermal growth factor and intestinal barrier function.
Mediat. Inflamm. 2016, 2016, 1927348. [CrossRef] [PubMed]
15. Berseth, C.L. Enhancement of intestinal growth in neonatal rats by epidermal growth factor in milk. Am. J.
Physiol. 1987, 253, G662–G665. [CrossRef] [PubMed]
16. Gavaldà-Navarro, A.; Hondares, E.; Giralt, M.; Mampel, T.; Iglesias, R.; Villarroya, F. Fibroblast growth factor
21 in breast milk controls neonatal intestine function. Sci. Rep. 2015, 5, 13717. [CrossRef] [PubMed]
17. Coursodon, C.F.; Dvorak, B. Epidermal growth factor and necrotizing enterocolitis. Curr. Opin. Pediatr. 2012,
24, 160–164. [CrossRef] [PubMed]
18. Kharitonenkov, A.; DiMarchi, R. Fibroblast growth factor 21 night watch: Advances and uncertainties in the
field. J. Intern. Med. 2017, 281, 233–246. [CrossRef] [PubMed]
19. Grases-Pintó, B.; Abril-Gil, M.; Rodríguez-Lagunas, M.J.; Castell, M.; Pérez-Cano, F.J.; Franch, À. Leptin and
adiponectin supplementation modifies mesenteric lymph node lymphocyte composition and functionality
in suckling rats. Br. J. Nutr. 2018, 119, 486–495. [CrossRef] [PubMed]
20. Rigo-Adrover, M.D.M.; Van Limpt, K.; Knipping, K.; Garssen, J.; Knol, J.; Costabile, A.; Franch, À.; Castell, M.;
Pérez-Cano, F.J. Preventive effect of a synbiotic combination of galacto- and fructooligosaccharides mixture
with Bifidobacterium breve M-16V in a model of multiple rotavirus infections. Front. Immunol. 2018, 9, 1318.
[CrossRef] [PubMed]
39
Nutrients 2018, 10, 1171
21. Maynard, A.A.; Dvorak, K.; Khailova, L.; Dobrenen, H.; Arganbright, K.M.; Halpern, M.D.; Kurundkar, A.R.;
Maheshwari, A.; Dvorak, B. Epidermal growth factor reduces autophagy in intestinal epithelium and in
the rat model of necrotizing enterocolitis. Am. J. Physiol. Liver Physiol. 2010, 299, G614–G622. [CrossRef]
[PubMed]
22. Camps-Bossacoma, M.; Pérez-Cano, F.J.; Franch, À.; Untersmayr, E.; Castell, M. Effect of a cocoa diet on the
small intestine and gut-associated lymphoid tissue composition in an oral sensitization model in rats. J. Nutr.
Biochem. 2017, 42, 182–193. [CrossRef] [PubMed]
23. Basha, S.; Surendran, N.; Pichichero, M. Immune responses in neonates. Expert Rev. Clin. Immunol. 2014, 10,
1171–1184. [CrossRef] [PubMed]
24. Hormi, K.; Lehy, T. Transforming growth factor-alpha in vivo stimulates epithelial cell proliferation in
digestive tissues of suckling rats. Gut 1996, 39, 532–538. [CrossRef] [PubMed]
25. Pollack, P.F.; Goda, T.; Colony, P.C.; Edmond, J.; Thornburg, W.; Korc, M.; Koldovský, O. Effects of enterally
fed epidermal growth factor on the small and large intestine of the suckling rat. Regul. Pept. 1987, 17,
121–132. [CrossRef]
26. Camporez, J.P.; Asrih, M.; Zhang, D.; Kahn, M.; Samuel, V.T.; Jurczak, M.J.; Jornayvaz, F.R. Hepatic insulin
resistance and increased hepatic glucose production in mice lacking FGF21. J. Endocrinol. 2015, 226, 207–217.
[CrossRef] [PubMed]
27. Pérez-Cano, F.J.; Franch, À.; Castellote, C.; Castell, M. The suckling rat as a model for immunonutrition
studies in early life. Clin. Dev. Immunol. 2012, 2012, 537310. [CrossRef] [PubMed]
28. Zhang, M.F. The Role of Milk Transforming Growth Factor-β (TGF-β) in the Development of the Infant Gut
and Gut Mucosal Immune System. Ph.D. Thesis, University of Adelaide, Adelaide, Australia, 2000.
29. Wills-Karp, M.; Luyimbazi, J.; Xu, X.; Schofield, B.; Neben, T.Y.; Karp, C.L.; Donaldson, D.D. Interleukin-13:
Central mediator of allergic asthma. Science 1998, 282, 2258–2261. [CrossRef] [PubMed]
30. Kubo, T.; Morita, H.; Sugita, K.; Akdis, C.A. Introduction to mechanisms of allergic diseases. In Middleton’s
Allergy Essentials; Holgate, S., Sheikh, A., Eds.; Elsevier: New York, NY, USA, 2017; pp. 1–27.
ISBN 978-0-323-37579-5.
31. Bao, K.; Reinhardt, R.L. The differential expression of IL-4 and IL-13 and its impact on type-2 immunity.
Cytokine 2015, 75, 25–37. [CrossRef] [PubMed]
32. Munblit, D.; Treneva, M.; Peroni, D.G.; Colicino, S.; Chow, L.Y.; Dissanayeke, S.; Pampura, A.; Boner, A.L.;
Geddes, D.T.; Boyle, R.J.; et al. Immune components in human milk are associated with early infant
immunological health outcomes: A prospective three-country analysis. Nutrients 2017, 9, 532. [CrossRef]
[PubMed]
33. Sakaguchi, K.; Koyanagi, A.; Kamachi, F.; Harauma, A.; Chiba, A.; Hisata, K.; Moriguchi, T.; Shimizu, T.;
Miyake, S. Breast-feeding regulates immune system development via transforming growth factor-β in mice
pups. Pediatr. Int. 2018, 60, 224–231. [CrossRef] [PubMed]
34. Dvorak, B.; Halpern, M.D.; Holubec, H.; Williams, C.S.; McWilliam, D.L.; Dominguez, J.A.; Stepankova, R.;
Payne, C.M.; McCuskey, R.S. Epidermal growth factor reduces the development of necrotizing enterocolitis
in a neonatal rat model. Am. J. Physiol. Gastrointest. Liver Physiol. 2002, 282, G156–G164. [CrossRef] [PubMed]
35. Weaver, L.T.; Walker, W.A. Epidermal growth factor and the developing human gut. Gastroenterology 1988,
94, 845–847. [CrossRef]
36. Wong, W.M.; Wright, N.A. Epidermal growth factor, epidermal growth factor receptors, intestinal growth,
and adaptation. JPEN J. Parenter. Enteral Nutr. 1999, 23, S83–S88. [CrossRef] [PubMed]
37. Halpern, M.D.; Dominguez, J.A.; Dvorakova, K.; Holubec, H.; Williams, C.S.; Meza, Y.G.; Ruth, M.C.;
Dvorak, B. Ileal cytokine dysregulation in experimental necrotizing enterocolitis is reduced by epidermal
growth factor. J. Pediatr. Gastroenterol. Nutr. 2003, 36, 126–133. [CrossRef] [PubMed]
38. Hanson, L.Å.; Korotkova, M. The role of breastfeeding in prevention of neonatal infection. Semin. Neonatol.
2002, 7, 275–281. [CrossRef] [PubMed]
39. Takeyama, K.; Dabbagh, K.; Lee, H.M.; Agustí, C.; Lausier, J.A.; Ueki, I.F.; Grattan, K.M.; Nadel, J.A.
Epidermal growth factor system regulates mucin production in airways. Proc. Natl. Acad. Sci. USA 1999, 96,
3081–3086. [CrossRef] [PubMed]
40
Nutrients 2018, 10, 1171
40. Yan, J.; Wang, J.; Huang, H.; Huang, Y.; Mi, T.; Zhang, C.; Zhang, L. Fibroblast growth factor 21 delayed
endothelial replicative senescence and protected cells from H2 O2 -induced premature senescence through
SIRT1. Am. J. Transl. Res. 2017, 9, 4492–4501. [PubMed]
41. Li, S.M.; Yu, Y.H.; Li, L.; Wang, W.F.; Li, D.S. Treatment of CIA mice with FGF21 down-regulates TH17-IL-17
axis. Inflammation 2016, 39, 309–319. [CrossRef] [PubMed]
© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
41
nutrients
Article
Effect of Milk Fermented with
Lactobacillus fermentum on the Inflammatory
Response in Mice
Lourdes Santiago-López 1 , Adrián Hernández-Mendoza 1 , Verónica Mata-Haro 2 ,
Belinda Vallejo-Córdoba 1 , Abraham Wall-Medrano 3 , Humberto Astiazarán-García 4 ,
María del Carmen Estrada-Montoya 1 and Aarón F. González-Córdova 1, *
1 Laboratorio de Química y Biotecnología de Productos Lácteos, Centro de Investigación en Alimentación y
Desarrollo A. C. (CIAD), Carretera a La Victoria Km. 0.6, Hermosillo, Sonora 83304, Mexico;
lulu140288@gmail.com (L.S.-L.); ahernandez@ciad.mx (A.H.-M.); vallejo@ciad.mx (B.V.-C.);
carmenes@ciad.mx (M.d.C.E.-M.)
2 Laboratorio de Microbiología e Inmunología, Centro de Investigación en Alimentación y Desarrollo
A. C. (CIAD), Carretera a La Victoria Km. 0.6, Hermosillo, Sonora 83304, Mexico; vmata@ciad.mx
3 Departamento de Ciencias Químico-Biológicas, Instituto de Ciencias Biomédicas, Universidad Autónoma de
Ciudad Juárez, Anillo Envolvente del PRONAF y Estocolmo s/n, Ciudad Juárez 32310, Chihuahua, Mexico;
awall@uacj.mx
4 Laboratorio de Patología Experimental, Centro de Investigación en Alimentación y Desarrollo A. C. (CIAD),
Carretera a la Victoria Km. 0.6, Hermosillo, Sonora 83304, Mexico; hastiazaran@ciad.mx
* Correspondence: aaronglz@ciad.mx; Tel./Fax: +52-662-289-2400
Abstract: Currently, the effect of fermented milk on the T-helper 17 response in inflammatory bowel
diseases (IBDs) is unknown. The aim of the present study was to evaluate the effect of milks
fermented with Lactobacillus fermentum on the Th1/Th17 response in a murine model of mild IBD.
Exopolysaccharide (EPS), lactic acid (LA), and total protein (TP) contents and bacterial concentration
were determined. Male C57Bl/6 mice intragastrically received either raw (FM) or pasteurized (PFM)
fermented milk before and during a dextran sulfate infusion protocol. Blood, spleen, and colon
samples were collected at Weeks 6 and 10. IL-6, IL-10, and TNFα were determined in serum, and IL-17,
IL-23, and IFNγ were determined in intestinal mucosa and serum. The FM groups did not differ in cell
concentration, LA, or TP content (p > 0.05); FM-J28 had the highest EPS content. Spleen weight and
colon length did not differ among the FM groups (p > 0.05). In the FM-J20 and PFM-J20 groups, IL-17
and IFNγ decreased, and the IL-10 concentration was enhanced (p < 0.05) at Week 6. IL-6, TNFα, IL-23,
and IFNγ did not differ in serum and mucosa (p > 0.05), and IL-17 was lowest in FM-J28 and FM-J20.
Therefore, FM appears to potentially play a role in decreasing the Th17 response. However, further
studies are needed to elucidate the FM-mediated anti-inflammatory mechanisms in IBD.
1. Introduction
Inflammatory bowel diseases (IBDs) are characterized by chronic and uncontrolled inflammation
in the intestinal mucosa. Different factors have been evidenced to affect the immune system at
the mucosal level [1]. For example, inflammatory mediators such as cytokines play an important
role in the adaptive immune response at the intestinal mucosal level. The modulation of biological
cellular functions may initiate downstream signaling pathways and mediate cellular proliferation and
differentiation [2]. In particular, Th1 and Th17 cells have been implicated in the development of IBD.
Th1 is characterized by the presence of interferon-γ (IFNγ) and Th17 by the presence of interleukin
(IL)-17, IL-21, and IL-22 [3]. Cytokines such as IL-6, TGF-β, and IL-23 promote the development of Th17
cells in IBD [4]. One study related Th1 and Th17 responses to the pathogenesis of IBD and suggested
that Th1 cells may enhance the production of Th17 cells. In this previous study, a higher concentration
of Th17 vs. Th1 cells was reported in a colitis model with CBirl TCR transgenic mice, which are
immunodominant susceptible to flagellin microbiota [5]. In contrast, another study examining the
cytokine profiles of Th1 and Th17 in a colitis model found that IL-4 and IL-10 levels were enhanced
while IL-17 levels were reduced [6].
In another study, the pathogenic action of IL-23 was demonstrated. In IL-23R-deficient mice,
reduced Reg3b protein expression in intestinal mucosa was shown to directly affect antimicrobial
activity. In addition, IL-23-dependent Reg3b triggers an influx of IL-22, regulating the number of
neutrophils in the lamina propia and restoring IL-22 secretion. This finding is important given the role
of IL-23 in neutralizing the Th17 response in IBD [7].
Moreover, the administration of probiotics was shown to possibly modulate the inflammatory
process through the Th1-Th2-Th17 response. Zheng et al. [8] reported that Bifidobacterium breve and
Lactobacillus rhamnosus reduce Th17 and increase the Th2 cell subset in human peripheral blood
mononuclear cells. In addition, the active components of probiotics were shown to be responsible
for enhancing the numbers of CD4 + FoxP3 + regulatory T cells in mesenteric lymph nodes and
for decreasing the cytokines tumor necrosis factor-α (TNFα), IFNγ, and IL-10 in Peyer’s patches
and the large intestine. In another study, probiotics were shown to play an important role in the
down-regulation of the nuclear factor kappa B (NFkB) pathway in RAW 264.7 cells to prevent TNFα
expression in a lipopolysaccharide-induced model [9].
Several additional studies have documented the regulation of the immune response in IBDs by
administering probiotics [10–14]. However, few studies have documented the effects of fermented
milk on the Th17 response, which regulates various inflammatory processes at the intestinal level.
In one case, Dahi-fermented milk containing a probiotic reduced myeloperoxidase (MPO) activity and
TNFα, IL-6, and IFNγ levels [14]. Meanwhile, milk with fermented Lactobacillus rhamnosus GG reduced
colonic inflammation and injury and stimulated the activation of epidermal growth factor receptor
(EGFR) and protein kinase B (Akt), which may be attributed to the release of p40 and p75 proteins
during fermentation [15].
These latter studies demonstrated the potential role of milk fermented with probiotics on the
inflammatory process. However, the effect of fermented milk on the Th1/Th17 response has not
been reported. The aim of the present study was to evaluate the effect of milk fermented with
Lactobacillus fermentum (J20 and J28) on the Th1/Th17 response in a murine model of inflammation.
43
Nutrients 2018, 10, 1039
by counts on plates of MRS agar. The LA and TP contents were determined by AOAC techniques
2000. The titratable acidity was expressed as percent LA titrated in 10 mL of fermented milk, using
NaOH (0.1 N) and phenolphtalein as the indicator. TP was quantified by the Kjeldahl method using a
nitrogen-to-protein conversion factor of 6.25.
44
Nutrients 2018, 10, 1039
Figure 1. Experimental design. FM + DSS, PFM + DSS, and AM experimental groups of mice
received daily treatments from Week 1 to Week 10. Mice received DSS in Week 2, 4, 6, and 9.
*DSS administration period
45
Nutrients 2018, 10, 1039
Table 1. Cell concentrations, total protein in fermented milk and metabolites derived from the
fermentation process.
46
Nutrients 2018, 10, 1039
Colon length did not differ significantly among all groups (p > 0.05) nor between Weeks 6 and 10,
which may indicate a low response to the inflammatory process during the study period.
The administration of DSS is known to reproducibly induce mild intestinal inflammation and
the development of ulcers in mice, increasing neutrophil counts in the intestine [24]. In this regard,
the capacity of Th17 cells to secrete IL-17 but not IFNγ or IL-4 has been described; these prior cells
play an important role in the mediation of the inflammation process and tissue destruction [25,26].
Therefore, it is important to know their functions and develop strategies that block their response at
local level, principally in IBDs [27].
The serum cytokine profiles at Week 6 showed a significant increase in IL-17 in the groups
subjected to DSS-induced inflammation with respect to the water group (p < 0.05) (Figure 3A).
Meanwhile, the concentration of IFNγ in the DSS group was enhanced compared to the water group at
Week 6, yet decreased by Week 10 (p < 0.05) (Figure 3B). The concentration of IFN-γ in serum samples
was enhanced at Week 6 for the DSS groups FM-J20 and PFM-J28. These values corresponded with an
enhancement in the Th1 response, as various reports have documented. The inflammation process,
as based on IL-17, was sufficient at Week 6; hence, by Week 10, the organisms were possibly able to
regulate the inflammation process. In our study, for example, the FM-J28 group at Week 6 showed
enhanced IL-17 but not IFN. This is a possible response mechanism to a low concentration of IFNγ.
Finally, IL-22 and IL-23 were not detected.
Figure 2. Effect of dextran sulfate sodium and fermented milk administration on spleen weight (A)
and colon length (B) at Weeks 6 (black) and 10 (grey). Bars with an uppercase letter indicate statistical
differences at Week 6 among treatments, and bars with a lowercase letter indicate statistical differences
at Week 10 according to the Tukey–Kramer test (p < 0.05). The values are the means ± SD (n = 5).
The groups administered with FM-J20 and PFM-J20 showed the lowest concentration of IL-17
(p < 0.05) at Weeks 6 and 10. FM-J20 and PFM-J20 differed in the presence of viable bacterial cells, yet
both FMs were able to decrease the concentration of IL-17, suggesting that different components from
FM are involved in this effect. However, no statistical difference was encountered at Week 10 with
respect to the DSS group (p > 0.05). Furthermore, IL-6 and TNFα did not show statistical differences
among treatments and the control (DSS and water) at Weeks 6 and 10 (p > 0.05) (Figure 3C,D).
However, the concentration of IL-10 cytokine did significantly differ (p < 0.05) (Figure 3E); the
groups administered with PFM-J20 and FM J28 presented the highest concentration at Weeks 6
and 10, respectively.
It was previously reported that the expression of the genes that encode for IL-6, IL-1β, IL-23A,
TGFβ, and STAT3 are involved in the differentiation of Th17 cells in a DSS-induced inflammation
model [28]. Furthermore, studies have demonstrated that Th17 cells require specific cytokines such as
47
Nutrients 2018, 10, 1039
IL-23, which mediates the expansion of IL-17. On the other hand, low levels of IFNγ and IFNα may
enhance the gene expression of IL-17 and IL-17-producing cells generated by IL-23 stimulation [29].
The findings reported in the present work are in agreement with those of other studies.
For example, the administration of VSL#3 probiotics decreased TNFα and IL-6 and increased IL-10
serum levels in a DSS-induced colitis model of inflammation [10]. However, the concentration of
inflammatory cytokines in our study is minor compared with other studies, which may indicate that
the response following FM intake is lower in our inflammation model. Different compounds derived
from FM are responsible for the anti-inflammatory effect [30,31]. For instance, milk fermented with
Lactobacillus rhamnosus GG (LGG) was found to reduce colonic inflammation and injury and to also
stimulate the activation of EGFR and Akt pathways while suppressing cytokine-induced apoptosis.
This effect was attributed to the soluble proteins p40 and p75 present in milk fermented with LGG [15].
Furthermore, the immunoprotective effect of probiotic Dahi containing Lactobacillus acidophilus LaVK2
and Bifidobacterium bifidum BbVK3 on DSS-induced ulcerative colitis in mice was demonstrated to
significantly reduce TNFα, IL-6, and IFNγ cytokines in colonic tissue [14].
48
Nutrients 2018, 10, 1039
Figure 3. Effects of DSS and fermented milk administration on: IL-17 (A); and IFNγ (B) in serum samples,
determined by ELISA; and on: IL-6 (C); TNFα (D); and IL-10 (E), determined by flow cytometry at
Weeks 6 and 10. Bars with an uppercase letter indicate statistical differences at Week 6 among treatments,
and bars with a lowercase letter indicate statistical differences at Week 10 (p < 0.05) according to the
Kruskal–Wallis test. The values correspond to medians ± interquartile ranges for n = 5.
Several reports have shown the different mechanisms of action involved in the anti-inflammatory
effect of probiotics, but few studies have demonstrated possible pathways at the cellular level.
49
Nutrients 2018, 10, 1039
One possible mechanism of action may be related with cell viability or cell wall components, which
can be internalized into M cells, interact with dendritic cells, and then down-regulate the production
of the pro-inflammatory cytokines IL-1β, IL-6, IL-17, TNFα, and IFNγ [32,33]. Furthermore, bioactive
peptides have shown anti-inflammatory activity by down-regulating LPS-induced cytokine production
in monocytes cells via the NF-kB pathway [34].
Moreover, LA may induce different inflammatory responses in LPS-stimulated RAW 264.7 cells.
Some studies have reported its association with metabolic acidosis, which frequently complicates
sepsis and septic shock and may be deleterious for cellular function. In this study, the AM group
showed no statistical difference (p > 0.05) with respect to the control group and FM groups; however,
it is important to carry out further studies on LA and its role in inflammatory processes.
In the determination of cytokines in samples of intestinal mucosal (Figure 4), FM-J20, FM-J28,
and PFM-J28 showed statistical differences (p < 0.05) among treatments with respect to the controls
(DSS and water). However, IFNγ levels were not statistically different (p > 0.05), and IL-23 was not
detected at the mucosal level. These results correspond with those reported in other studies wherein
the presence of IFNγ inhibited IL-17 [5]. IL-23 is important for the expansion, stabilization, and
conditioning of the Th17 response; hence, the IL-23/17 axis may be considered as a hallmark and an
attractive probiotic therapeutic target in IBD [32]. The study by Jadhav et al. [14] showed the lowest
concentrations of TNFα, IL-6, and IFNγ in samples of colonic tissue from mice administered with Dahi
(fermented dairy) + DSS as well as a group administered probiotic + DSS.
This effect has also been observed in mice following the administration of milk fermented with
Bifidobacterium strains. In particular, these mice showed a lower concentration of TNFα in culture
supernatants, while IL-10 increased in the FM group and also reduced the histological score compared
to treatments with saline and unfermented milk [35]. A possible mechanism for the anti-inflammatory
effect may be related to the inhibition or modulation of the expression of several genes such as NF-kB,
RELB, and TNFα, which have been reported as active during inflammation processes [36].
Figure 4. The effect of DSS and fermented milk administration on IL-17 (A) and IFNγ (B) determined
by ELISA of colonic mucosa at Week 10. Different letters show statistical differences (p < 0.05) among
treatments according to the Kruskal–Wallis test. The values are medians ± interquartile ranges (n = 5).
50
Nutrients 2018, 10, 1039
Histological Analysis
The DSS inflammation process is characterized by histological findings such as edema, infiltration
of inflammatory cells into both the mucosa and submucosa, and destruction of epithelial cells [28].
Our results indicated that neither the administration of DSS or fermented milk affected the structure of
mucosa or submucosa (Figure 5). A low number of inflammatory cells infiltrated (Figure 5B) but did not
result in mucosal injury. Therefore, the histological analysis of samples confirmed that the inflammation
process was not extensive; this finding may suggest that the presence of pro-inflammatory cytokines
did not cause extensive changes. On the other hand, some studies have suggested that invasion of
inflammatory cells into the mucosa produces increased concentrations of inflammatory cytokines such
as TNFα, which then induce the expression of genes associated with inflammation [10]. In addition,
DSS may induce the expression of COX-2, an enzyme responsible for the formation of prostanoids [38]
that has been shown to be specifically induced in epithelial cells under IBD conditions [39].
Figure 5. Histological analysis of cross sections of colon tissue samples stained with hematoxylin–eosin
(10×): (A) water control group; (B) DSS group; (C) AM + DSS group; (D) FM-J20 + DSS group;
(E) PFM-J20 + DSS group; (F) FM-J28 + DSS group; and (G) PFM-J28 + DSS group.
These findings show that FM administration may prevent intestinal inflammation by stabilizing
mucosal immunity through different components or metabolites derived during the fermentation
process [13]. In particular, the present results show that the administration of FM could mediate the
Th1/Th17 response, but more studies are required to determine the possible metabolites involved
and the activation pathways. The interaction of derivative compounds from fermentation, such as
LA and EPS, as well as the presence of bacteria could result in an anti-inflammatory response that
stimulates IL-10 production and inhibits TNFα despite the interaction of TLR2 and TLR4 and activation
of NF-kB. For example, in one previous study, the presence of LA and Lactobacillus casei Shirota culture
supernatants suppressed phosphorylation and degradation of I-kB-α [40].
4. Conclusions
In the present study, several strategies to reduce inflammation in an IBD model employing
milk fermented with Lactobacillus strains were evaluated. The administration of milk fermented
51
Nutrients 2018, 10, 1039
with Lactobacillus fermentum possibly decreased the inflammatory response at Week 6 because of
the metabolites or cell components present in this product. However, further studies are needed to
determine the modulation of Th1/Th17 by fermented milk. The findings in the present study show the
potential regulatory effect of FM on the inflammatory process, although this effect was minor, possibly
as a result of irregularity in the inflammatory process. Future studies are needed to establish an
adequate model of inflammation that allows the Th17 response to be evaluated considering the other
biomarkers involved in this response, including chemokines, transcription factors, and metabolites
released during fermentation, as well as cell differentiation, which may all be responsible for promoting
an anti-inflammatory response.
Author Contributions: A.F.G.-C., A.H.-M., V.M.-H., B.V.-C., A.W.-M., H.A.-G. and L.S.-L. designed the study.
V.M.-H. and L.S.-L. contributed to the flow cytometer analysis of the samples. L.S.-L. wrote the manuscript.
V.M.-H., A.H.-M. and A.F.G.-C. revised the manuscript. A.W.-M. contributed to the histological analysis. H.A.-G.,
B.V.-C. and M.d.C.E.-M. provided valuable scientific knowledge and advisory throughout the study.
Funding: This study was supported by the Mexican Council of Science and Technology (CONACyT; Mexico City)
research project 240338 CONACyT.
Acknowledgments: The authors would like to thank Alejandro Santos-Espinosa, Alejandro Epigmenio-Chavez,
Lilia María Beltrán-Barrientos, and Miguel Angel Rendón-Rosales for their technical assistance in this study.
The authors would also like to thank the Mexican Council of Science and Technology (CONACyT) for the graduate
scholarship provided to L. Santiago-López. Histological analyses were performed with equipment acquired with
funding from the Comprehensive Institutional Strengthening Program (PIFI) 2007-2008, 2909 5001-004-09
Conflicts of Interest: The authors declare that they have no conflicts of interest regarding the publication of
this paper.
References
1. Chami, B.; Yeung, A.W.S.; van Vreden, C.; King, N.J.C.; Bao, S. The role of CXCR3 in DSS-induced colitis.
PLoS ONE 2014, 9, e101622. [CrossRef] [PubMed]
2. O’Shea, J.J.; Murray, P.J. Cytokine signaling modules in inflammatory responses. Immunity 2008, 28, 477–487.
[CrossRef] [PubMed]
3. Ito, R.; Kita, M.; Shin-Ya, M.; Kishida, T.; Urano, A.; Takada, R.; Sakagami, J.; Imanishi, J.; Iwakura, Y.;
Okanoue, T.; et al. Involvement of IL-17A in the pathogenesis of DSS-induced colitis in mice. Biochem. Biophys.
Res. Commun. 2008, 377, 12–16. [CrossRef] [PubMed]
4. Weaver, C.T.; Elson, C.O.; Fouser, L.A.; Kolls, J.K. The Th17 Pathway and inflammatory diseases of the
intestines, lungs, and skin. Annu. Rev. Pathol. Mech. Dis. 2013, 8, 477–512. [CrossRef] [PubMed]
5. Feng, T.; Qin, H.; Wang, L.; Benveniste, E.N.; Elson, C.O.; Cong, Y. Th17 cells induce colitis and promote Th1
cell responses through IL-17 induction of innate IL-12 and IL-23 production. J. Immunol. 2011, 186, 6313–6318.
[CrossRef] [PubMed]
6. Alex, P.; Zachos, N.C.; Nguyen, T.; Gonzales, L.; Chen, T.E.; Conklin, L.S.; Centola, M.; Li, X. Distinct cytokines
patterns identified from multiplex profiles of murine DSS and TNBS-Induced Colitis. Inflamm. Bowel Dis.
2009, 15, 341–352. [CrossRef] [PubMed]
7. Aden, K.; Rehman, A.; Falk-Paulsen, M.; Secher, T.; Kuiper, J.; Tran, F.; Pfeuffer, S.; Sheibani-Tezerji, R.;
Breuer, A.; Luzius, A.; et al. Epithelial IL-23R signaling licenses protective IL-22 responses in intestinal
inflammation. Cell Rep. 2016, 16, 2208–2218. [CrossRef] [PubMed]
8. Zheng, B.; van Bergenhenegouwen, J.; Overbeek, S.; van de Kant, H.J.G.; Garssen, J.; Folkerts, G.; Vos, P.;
Morgan, M.E.; Kraneveld, A.D. Bifidobacterium breve attenuates murine dextran sodium sulfate-induced
colitis and increases regulatory T cell responses. PLoS ONE 2014, 9, e95441. [CrossRef] [PubMed]
9. Zakostelska, Z.; Kverka, M.; Klimesova, K.; Rossmann, P.; Mrazek, J.; Kopecny, J.; Hornova, M.; Srutkova, D.;
Hudcovic, T.; Ridl, J.; et al. Lysate of probiotic Lactobacillus casei DN-114 001 ameliorates colitis by
strengthening the gut barrier function and changing the gut microenvironment. PLoS ONE 2011, 6, e27961.
[CrossRef] [PubMed]
10. Dai, C.; Zheng, C.Q.; Meng, F.J.; Zhou, Z.; Sang, L.X.; Jiang, M. VSL#3 probiotics exerts the anti-inflammatory
activity via PI3k/Akt and NF-κB pathway in rat model of DSS-induced colitis. Mol. Cell. Biochem. 2013,
374, 1–11. [PubMed]
52
Nutrients 2018, 10, 1039
11. Chiba, Y.; Shida, K.; Nagata, S.; Wada, M.; Bian, L.; Wang, C.; Shimizu, T.; Yamashiro, Y.; Kiyoshima-Shibata, J.;
Nanno, M.; et al. Well-controlled proinflammatory cytokine responses of Peyer’s patch cells to probiotic
Lactobacillus casei. Immunology 2010, 130, 352–362. [CrossRef] [PubMed]
12. Herías, M.V.; Koninkx, J.F.; Vos, J.G.; Huis in’t Veld, J.H.; van Dijk, J.E. Probiotic effects of Lactobacillus casei
on DSS-induced ulcerative colitis in mice. Int. J. Food Microbiol. 2005, 103, 143–155. [CrossRef] [PubMed]
13. Imaoka, A.; Umesaki, Y. Rationale for Using of Bifidobacterium probiotic strains-fermented milk against colitis
based on animal experiments and clinical trials. Probiotics Antimicrob. Proteins 2008, 1, 8–14. [CrossRef]
[PubMed]
14. Jadhav, S.R.; Shandilya, U.K.; Kansal, V.K. Immunoprotective effect of probiotic Dahi Containing
Lactobacillus acidophilus and Bifidobacterium bifidum on dextran sodium sulfate-induced ulcerative colitis
in mice. Probiotics Antimicrob. Proteins 2012, 4, 21–26. [CrossRef] [PubMed]
15. Yoda, K.; Miyazawa, K.; Hosoda, M.; Hiramatsu, M.; Yan, F.; He, F. Lactobacillus GG-fermented milk prevents
DSS-induced colitis and regulates intestinal epithelial homeostasis through activation of epidermal growth
factor receptor. Eur. J. Nutr. 2014, 53, 105–115. [CrossRef] [PubMed]
16. Tallon, R.; Bressollier, P.; Urdaci, M.C. Isolation and characterization of two exopolysaccharides produced by
Lactobacillus plantarum EP56. Res. Microbiol. 2003, 154, 705–712. [CrossRef] [PubMed]
17. Nanda-Kumar, N.S.; Balamurugan, R.; Jayakanthan, K.; Pulimood, A.; Pugazhendhi, S.; Ramakrishna, B.S.
Probiotic administration alters the gut flora and attenuates colitis in mice administered dextran sodium
sulfate. J. Gastroenterol. Hepatol. 2008, 23, 1834–1839. [CrossRef] [PubMed]
18. Taverniti, V.; Guglielmetti, S. The immunomodulatory properties of probiotic microorganisms beyond their
viability (ghost probiotics: Proposal of paraprobiotic concept). Genes Nutr. 2011, 6, 261–274. [CrossRef] [PubMed]
19. Hidalgo-Cantabrana, C.; Algieri, F.; Rodriguez-Nogales, A. Effect of a Ropy Bifidobacterium animalis subsp.
lactis Strain orally administered on DSS-induced colitis mice Model. Front. Microbiol. 2016, 7, 868. [CrossRef]
[PubMed]
20. Kellum, J.A.; Song, M.; Li, J. Lactic and hydrochloric acids induce different patterns of inflammatory
response in LPS-stimulated RAW 264.7 cells. Am. J. Physiol. Regul. Intregr. Comp. Physiol. 2004, 286, 686–692.
[CrossRef] [PubMed]
21. Degeest, B.; Janssens, B.; De Vuyst, L. Exopolysaccharide (EPS) biosynthesis by Lactobacillus sakei 0-1:
Production kinetics, enzyme activities and EPS yields. J. Appl. Microbiol. 2001, 91, 470–477. [CrossRef]
[PubMed]
22. Morgan, M.E.; Zheng, B.; Koelink, P.J.; van de Kant, H.J.G.; Haazen, L.C.; van Roest, M.; Garssen, J.;
Folkerts, G.; Kraneveld, A.D. New Perspective on Dextran Sodium sulfate colitis: Antigen-specific T cell
development during intestinal inflammation. PLoS ONE 2013, 8, e69936. [CrossRef] [PubMed]
23. Bronte, V.; Pittet, M. The spleen in local and systemic regulation of immunity. Immunity 2014, 39, 806–818.
[CrossRef] [PubMed]
24. Ohtsuka, Y.; Sanderson, I. Dextran sulfate Sodium-Induced inflammation is enhanced by intestinal epithelial
cell chemokine expression in mice. Pediatr. Res. 2003, 53, 143–147. [PubMed]
25. Fischer, A. Human immunodeficiency: Connecting STATA3, Th17 and human mucosal immunity.
Immunol. Cell Biol. 2008, 86, 549–551. [CrossRef] [PubMed]
26. Tesmer, L.; Lundy, S.; Sarkar, S.; Fox, D. Th17 cells in human disease. Immunol. Rev. 2008, 223, 87–113.
[CrossRef] [PubMed]
27. Crome, S.Q.; Wang, A.Y.; Levings, M.K. Translational mini-review series on Th17 cells: Function and
regulation of human T helper 17 cells in health and disease. Clin. Exp. Immunol. 2010, 159, 109–119.
[CrossRef] [PubMed]
28. Ogawa, A.; Andoh, A.; Araki, Y.; Bamba, T.; Fujiyama, Y. Neutralization of interleukin-17 aggravates dextran
sulfate sodium-induced colitis in mice. Clin. Immunol. 2004, 110, 55–62. [CrossRef] [PubMed]
29. Liu, Z.J.; Yadav, P.K.; Su, J.L.; Wang, J.S.; Fei, K. Potential role of Th17 cells in the pathogenesis of inflammatory
bowel disease. World J. Gastroenterol. 2009, 15, 5784–5788. [CrossRef] [PubMed]
30. Granier, A.; Goulet, O.; Hoarau, C. Fermentation products: Immunological effects on human and animal
models. Pediatr. Res. 2013, 74, 238–244. [CrossRef] [PubMed]
31. Bordoni, A.; Danesi, F.; Dardevet, D.; Dupont, D.; Fernandez, A.S.; Gille, D.; Nunes, C.; Pinto, P.; Re, R.;
Rémond, D.; et al. Dairy products and inflammation: A review of the clinical evidence. Crit. Rev. Food
Sci. Nutr. 2017, 57, 2497–2525. [CrossRef] [PubMed]
53
Nutrients 2018, 10, 1039
32. Owaga, E.; Hsieh, R.-H.; Mugendi, B.; Masuku, S.; Shinh, C.-K.; Chang, J.-S. Th17 cells as potential probiotic
therapeutic targets in inflammatory bowel diseases. Int. J. Mol. Sci. 2015, 16, 20841–20858. [CrossRef]
[PubMed]
33. Lee, H.S.; Han, S.Y.; Bae, E.A.; Huh, C.H.S.; Ahn, Y.T.; Lee, J.H.K.; Kim, D.H. Lactic acid bacteria inhibit
proinflammatory cytokine expression and bacterial glycosaminoglycan degradation activity in dextran
sulfate sodium-induced colitic mice. Int. Immunopharmacol. 2008, 8, 574–580. [CrossRef] [PubMed]
34. Håversen, L.; Ohlsson, B.G.; Hahn-Zoric, M.; Hanson, L.Å.; Mattsby-Baltzer, I. Lactoferrin down-regulates the
LPS-induced cytokine production in monocytic cells via NF-κB. Cell. Immunol. 2002, 220, 83–95. [CrossRef]
35. Matsumoto, S.; Watanabe, N.; Imaka, A.; Okabe, Y. Preventive effects of Bifidobacterium and
Lactobacillus-fermented milk on the development of inflammatory bowel disease in senescence-accelerated
mouse P1/Yit strain mice. Digestion 2001, 64, 92–99. [CrossRef] [PubMed]
36. Haileselassie, Y.; Navis, M.; Vu, N.; Qazi, K.R.; Rethi, B.; Sverremark-Ekström, E. Postbiotic modulation
of retinoic acid imprinted mucosal-like dendritic cells by probiotic Lactobacillus reuteri 17938 in vitro.
Front. Immunol. 2016, 7, 96. [CrossRef] [PubMed]
37. Mirpuri, J.; Sotnikov, I.; Denning, T.L.; Yarovinsky, F.; Parkos, C.A.; Denning, P.W.; Louis, N.A.
Lactobacillus rhamnosus (LGG) regulates IL-10 signaling in the developing murine colon through upregulation
of the IL-10R2 receptor subunit. PLoS ONE 2012, 7, e51955. [CrossRef] [PubMed]
38. Krieglstein, C.F.; Cerwinka, W.H.; Laroux, F.S.; Salter, J.W.; Russell, J.M.; Schuermann, G.; Grisham, M.B.;
Ross, C.R.; Granger, D.N. Regulation of murine intestinal inflammation by reactive metabolites of oxygen
and nitrogen: Divergent roles of superoxide and nitric oxide. J. Exp. Med. 2001, 194, 1207–1218. [CrossRef]
[PubMed]
39. Menchen, L.; Colon, A.L.; Madrigal, J.L.; Beltran, L.; Botella, S.; Lizasoain, I.; Leza, J.C.; Moro, M.A.;
Menchen, P.; Cos, E.; et al. Activity of inducible and neuronal nitric oxide synthases in colonic mucosa
predicts progression of ulcerative colitis. Am. J. Gastroenterol. 2004, 99, 1756–1764. [CrossRef] [PubMed]
40. Watanabe, T.; Nishio, H.; Tanigawa, T.; Yamagami, H.; Okazaki, H.; Watanabe, K.; Tominaga, K.; Fujiwara, Y.;
Oshitani, N.; Asahara, T.; et al. Probiotic Lactobacillus casei strain Shirota prevents indomethacin-induced small
intestinal injury: Involvement of lactic acid. Am. J. Physiol. Gastrointest. Liver Physiol. 2009, 297, G506–G513.
[CrossRef] [PubMed]
© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
54
nutrients
Review
Infant Complementary Feeding of Prebiotics for the
Microbiome and Immunity
Starin McKeen 1,2,3 , Wayne Young 1,2,3 , Jane Mullaney 1,2,3 , Karl Fraser 1,2,3 ,
Warren C. McNabb 2,3 and Nicole C. Roy 1,2,3, *
1 AgResearch, Food Nutrition & Health, Grasslands Research Centre, Private Bag 11008, Palmerston north
4442, New Zealand; Starin.Mckeen@agresearch.co.nz (S.M.); Wayne.Young@agresearch.co.nz (W.Y.);
Jane.Mullaney@agresearch.co.nz (J.M.); Karl.Fraser@agresearch.co.nz (K.F.)
2 Riddet Institute, Massey University, Private Bag 11222, Palmerston North 4442, New Zealand;
W.McNabb@massey.ac.nz
3 High-Value Nutrition National Science Challenge, Auckland, New Zealand
* Correspondence: Nicole.Roy@agresearch.co.nz; Tel.: +64-6-3518-1101
Abstract: Complementary feeding transitions infants from a milk-based diet to solid foods, providing
essential nutrients to the infant and the developing gut microbiome while influencing immune
development. Some of the earliest microbial colonisers readily ferment select oligosaccharides,
influencing the ongoing establishment of the microbiome. Non-digestible oligosaccharides
in prebiotic-supplemented formula and human milk oligosaccharides promote commensal
immune-modulating bacteria such as Bifidobacterium, which decrease in abundance during weaning.
Incorporating complex, bifidogenic, non-digestible carbohydrates during the transition to solid foods
may present an opportunity to feed commensal bacteria and promote balanced concentrations of
beneficial short chain fatty acid concentrations and vitamins that support gut barrier maturation and
immunity throughout the complementary feeding window.
1. Introduction
The strategic introduction of prebiotic compounds during weaning presents an opportunity to
promote infant health and to support development via balanced co-maturation of the gut microbiome
and host. Between 4 and 6 months of age, nutrient demands of growing infants surpass what is
provided by breastmilk or formula alone [1–4]. Complementary foods accompany and gradually
replace breastmilk and formula throughout the weaning period, providing essential nutrients to the
developing digestive system and modulating microbial colonisation [1,5–8]. The young immune
system is influenced by the gut microbiome and supported by metabolites produced during the
microbial fermentation of prebiotic compounds, leading to a tolerance for commensal microbes
and specific responses to pathogens [9–15]. Prebiotic compounds in breastmilk and supplemented
formulas promote commensal immune-modulating bacteria, such as Bifidobacterium, and beneficial
metabolites, such as short chain fatty acids (SCFAs) and vitamins [16–21]. Introducing non-digestible
starches through complementary foods may present an opportunity to promote commensal bacteria
and support microbial production of beneficial metabolites throughout the complementary feeding
window, with lasting effects on health [22–24].
Prior to weaning, the healthy infant gut microbiome is shaped by maternal factors, such as mode
of birth, environment, and first foods: breastmilk and infant formula [10,25–33]. The establishment of
microbial species changes dramatically throughout the first 2–3 years of life before stabilising at an
adult-like composition [7]. While individual variations in taxonomic composition persist, analogous
genes consistently and predictably fill similar functional and metabolic niches as new foods are
introduced and formula or breastfeeding ceases [7]. Commensal species that colonise the immature
gut modulate gene expression of epithelial and immune cells and, in turn, are regulated by adaptive
and innate immune responses in the mucosal immune system [14,26,31,34–42].
Breastmilk and some types of prebiotic-supplemented formulas provide non-digestible
oligosaccharides (NDOs) to the gut microbiome, which exert a strong influence on the microbial
composition and metabolism [43]. The introduction of starchy foods such as cereals, porridges,
and pureed tubers is common practice due to the neutral tastes, smooth textures, and ease of
swallowing as oral coordination develops [44]. The role of these starches in the community dynamics
of the immature and unstable infant microbiome remains unknown.
Based on investigations into human milk oligosaccharides (HMOs) and NDOs, prebiotic whole
foods may support immunity and immune development through a variety of direct and indirect
mechanisms. While poorly characterised compared to oligosaccharides, starches may act as receptor
analogues to pathogens, reducing the quantity of enteric pathogens that reach the gut epithelium and
subsequent infection [45]. Starches also promote populations of bacteria of which some strains directly
interact with immunomodulatory factors in the gut mucosa [46]. These and other commensal bacteria
also ferment starches into metabolites such as SCFAs and vitamins, which have known benefits to gut
barrier integrity, immune-regulation, and immune response [47].
This review summarises the current body of knowledge on the complementary feeding of
prebiotic starches for the microbiome with a focus on the interactions of commensal species, microbial
metabolites, and the development of the gut barrier and immune system.
56
Nutrients 2019, 11, 364
Figure 1. The percent of nutrient requirements based on the recommended daily intakes (RDIs) [49]
that are met via average daily breastmilk consumption (750 mL from 0–6 months and 800 mL from
7–12 months) [50].
Timing of the introduction of solid foods has been investigated in both low- and high-income
countries. Delaying solids until 6 months of age was previously thought to be associated with
lower body mass index in high-income countries and with lower rates of allergy and decreased
water-borne diarrheal disease in low- and middle-income countries [51,52]. However, recent studies
in larger cohorts have challenged this assertion, proposing that individual oral maturation, nutrient
requirements, and environmental disease burden should determine when to introduce solids [4].
Results from the PIAMA (Prevention and Incidence of Asthma and Mite Allergy) cohort in the
Netherlands suggest that a short duration of breastfeeding (4 months or less) is associated with an
increased risk for being overweight during childhood rather than the early introduction of solid foods,
and the risk is not different between breastfed and formula-fed infants [53]. However, this study does
not report on the types of solid foods that were introduced, the duration of the overlap of breastfeeding
and solid feeding, or the potential mechanisms of metabolic programming.
In addition to nutritional provisions, breastmilk also provides non-nutritive and
immune-modulatory factors that impart significant benefits, even in partial concentrations or
shorter durations [12]. The health promoting properties of breastmilk include varying levels and
types of carbohydrates, non-digestible HMOs, immunoglobulins (IgG, IgM, and isoforms of sIgA),
amino acids, polyunsaturated fatty acids, monoglycerides, leuric acid, linoleic acid, cytokines,
57
Nutrients 2019, 11, 364
chemokines, soluble receptors, antibacterial proteins/peptides, and intact immune cells that are
governed by maternal Lewis blood type, secretor status, and phase of lactation [54]. HMOs have
received significant attention in infant nutrition for their ability to influence a variety of gut functions:
epithelial integrity, mucosal integrity, susceptibility to pathogenic infection, microbial community
structure, SCFA production, and vitamin synthesis. Over 2000 distinct HMO structures (Figure 2a)
have been identified, with significant variation between individuals and phase of lactation, but a 9:1
ratio of galactooligosaccharides (GOS):fructoligosaccharides (FOS) is typical [55,56]. Infant formulas
are continuing to develop based on an increasing understanding of the roles of each of these factors in
microbiome maturation, brain development, and immunity. Synthetic and plant-derived GOS, FOS
(Figure 2b), inulin, pectin, and β-glucans, either alone or in comparable ratios, are well characterised
and have been primary targets of infant formula additive research and product development [56,57].
Staged and follow-on formulas that vary in composition according to the recommended daily
allowances and the introduction of complementary foods are also increasingly recommended [2].
Figure 2. (a) The core structures of human milk oligosaccharides (HMOs), common modification
pathways, and an example of a complex HMO, connected by β1-3 and β 1-6 linkages that are resistant
to enzymatic cleavage by human-derived enzymes. (b) The structure of galactooligosaccharide (long
chain) and fructooligosaccharide (short chain), which are common prebiotic molecules in supplemented
infant formulas: β1-2, β1-4, and β1-6 linkages are resistant to enzymatic cleavage by human derived
enzymes. (c) A model of dietary starch, characterized by glucose molecules connected by α1-6 linkages
in a complex higher structure, which contributes to incomplete enzymatic cleavage by human enzymes.
Infant digestive systems are uniquely suited to digest macronutrients provided by breastmilk.
The intestinal epithelium of neonates has narrow villi and small crypts (Figure 3), which duplicate
and expand with age, a process which is influenced by components in breastmilk and host-microbe
interactions [58]. The expansion of the epithelial surface during weaning is necessary to accommodate
58
Nutrients 2019, 11, 364
the increasing nutrient load, but dysregulation of this process can lead to hyperplastic crypts, blunted
villi, inflammatory responses in the mucosa, and subsequent malabsorption of nutrients [59].
The enzymatic dynamics of infant digestion are poorly characterised due to wide variations
between individuals over time and infrequent investigations with replicated results [60]. Lactose, fatty
acids, and proteins are the most abundant macronutrients in milk, which are absorbed and utilised
predominantly in the small intestine [20]. Lipase and trypsin (lipid and protein digestive enzymes) are
present in concentrations comparable to adults and are sufficiently active at the less extreme pH (3.2) of
the infant gut. However, amylase secretion and activity are distinct in infants. Compared to lipid and
protein digestion, the ability to digest carbohydrates is limited to simple carbohydrates such as lactose
and sucrose, rather than complex carbohydrates, until weaning. At weaning, salivary α-amylase and
pancreatic α-amylase are present at reduced concentrations compared to that of adults [61]. However,
glucoamylase (also referred to as amyloglucosidase), a brush border enzyme in the small intestine
capable of cleaving α1,4-glycosidic bonds, is produced at 100–150% of adult concentrations at birth,
which may compensate for the otherwise minimal starch hydrolysis [62,63]. Non-digestible structures
such as HMOs and non-digestible carbohydrates (NDCs) resist complete enzymatic degradation
and pass to the large intestine where they become available as a nutrient source for the enteric
microbiota [64]. Breastfed infants also receive varying amounts of maternal amylase, as well as small
concentrations of up to 50 other digestive enzymes, through breastmilk [65]. During weaning, infants
that continue to consume breastmilk may have increased capacity to digest dietary starches compared
to those receiving formulas, but this and the subsequent interactions with enteric microbes has yet to
be investigated.
59
Nutrients 2019, 11, 364
Mucus Membrane
At the luminal surface of the enteric epithelium, the mucus layer provides a structural and
functional barrier that provides lubrication and separates the microbiota from epithelial cells while
allowing for the transport of nutrients and metabolites. Mucus is a complex heterogeneous suspension
matrix with high concentrations of high molecular weight glycoproteins called mucins, which are
secreted by goblet cells, and contains antimicrobial peptides, such as defensins [41]. Different types of
mucins have different roles in the lumen: secreted mucins form the mucus layer over the epithelium,
transmembrane mucins appear to be involved in signaling pathways, and some species of bacteria rely
on mucins as an energy source [40].
Several bacterial products, including lipopolysaccharides and flagellin on gram-negative bacteria
and lipoteichoic acids on gram positive bacteria, have been found to upregulate the mucin gene
expression and to stimulate mucin secretion [74]. Some probiotics, such as specific strains of
Bifidobacterium and Lactobacillus, successfully adhere to mucins and reach epithelial surfaces using
non-flagellar appendages called tight adherence pili, which influence immune responses [75,76].
This contributes to differences between the discarded microbiome identified in faecal collections and
the microbiome in the mucosa and at the epithelial surface [77,78]. Probiotic treatment, particularly
with Lactobacillus, has been shown to increase mucin and defensin secretion in murine models and
in vitro cell monolayers [79].
Prebiotics influence the composition of mucus by increasing the concentration of glycans [24],
decreasing the luminal pH, and increasing mucin glycosylation and sulphation [80], which protects
mucins from being degraded by host proteases and bacterial glycosidases (Figure 3) [81]. Mucus,
specifically secreted MUC-2, has also demonstrated immune-regulatory signals by interfering with the
expression of inflammatory cytokines but not tolerogenic cytokines by inhibiting gene transcription
through nuclear factor NF-κB (the nuclear factor kappa-light-chain-enhancer of activated B cells)
in dendritic cells (DCs) [82]. The mucus layer plays a significant role in microbial signaling,
cross-feeding, microbe-host interactions, and enteric immunity but can only be partially simulated in
in vitro experiments.
4. Establishment of the Microbiome and Immune System in the First Year of Life
Microbial community composition during the first year of life is dynamic, unstable, and susceptible
to perturbations [6,83]. The gut is the largest immune organ in the human body, containing
approximately 65% of immunologic tissues and up to 80% of the immunoglobulin-producing tissues
of the body [84]. During gestation, the foetal immune system is downregulated, making neonates
particularly susceptible to infection and aberrant immune responses. The epithelial barrier, mucosa,
and environmental conditions, such as pH, provide the majority of protection against pathogens in
the neonatal period (Figure 3) [85,86]. Healthy immune development in infants is characterised by
a transition from innate type 1 immunity, dominated by non-specific macrophages and neutrophils,
to adaptive type 2 immunity, characterised by specific T cells and B cells, which is fundamental
to the establishment of tolerance: the ability to distinguish between beneficial commensal bacteria
and harmful pathogens, leading to the appropriate scale and duration of responses to actual threats
(Figure 3) [87]. Spatial and temporal interactions between the microbiome, microbial metabolites,
and gut epithelial cells in the lumen, on the surface of epithelial cells, and in the interior components
of the gut-associated lymphoid tissue (GALT), such as DCs, modulate balanced immune development,
immune response, homeostasis, and disease (Figure 3) [88].
60
Nutrients 2019, 11, 364
Figure 3. A schematic of multiple mechanisms by which prebiotics modulate immune and gut
development. A. Prebiotics bind to pathogens as receptor analogues, preventing adhesion to the
epithelial surface and subsequent infection. B. Prebiotics promote populations of commensal microbes,
which outcompete pathogens for resources D, reducing infections. C. Prebiotics act directly upon the
epithelium promoting the mRNA transcription of proteins involved in barrier integrity. E. Commensal
microbes produce metabolites, such as short chain fatty acids (SCFAs), that decrease the lumen pH
and increase mucus F, increase TJ proteins and crypt and villi development G, and serve as an
energy source for enterocytes that form the epithelium H. In infants, the immature gut is susceptible
to allergy and pathogen translocation I through leaky gut barrier. J. Non-specific immune factors,
such as macrophages and neutrophils attack commensals and pathogens alike in poorly regulated
inflammatory responses. During immune development, dendritic cells K sample commensal microbes,
through Toll-Like Receptor (TLR) recognition, allowing for antigen specific immunoglobin production
L and promoting the differentiation of T and B cells M, resulting in improved tolerance to commensals
and targeted response to pathogens N.
61
Nutrients 2019, 11, 364
have been implicated in gut homeostasis and inflammatory responses characteristic of food allergies,
intestinal inflammation, and infections when poorly regulated [92]. Insufficient TLR exposure to
commensals, as found through antibiotic-mediated dysbiosis in murine models, is also correlated with
increased susceptibility to viral infections [93]. Infant TLR responses to commensal microbes differ
from responses in adults, demonstrating the impaired production of inflammatory mediators and
heightened production of inflammatory cytokines, such as IL-10 [85,87].
TLRs are susceptible to modulation by dietary starches in in vitro models. Different starch
structures bind differentially to TLRs, activating NF-κB, and activator proteins (AP-1), but the strong
immune-stimulating effects may also be attenuated by starch-exposed intestinal epithelial cells [94].
B2→1 fructans and High Maize 260 mainly stimulate TLR2, whereas Novelose 330 binds to TLR2
and TLR5 [95]. High Maize 260, which has a smaller average particle size of 12.8 μm, smooth surface,
and high degree of molecular order was found to have a stronger regulatory effect on epithelial cells
than Novelose 330, which has a larger average particle size of 46.6 μm and consists of destroyed and
convoluted granules due to the retrogradation process. Despite the attenuating activity, TLRs continue
to produce Th1 cytokines [94]. High-maize 260 is also more effective than inulin and sugar pectin in
reducing chemokine release in response to Sphingomonas paucimobilis infections in vitro [96]. In vivo,
the mucosal matrix is expected to drastically alter the exposure of epithelial cells to starch structures,
limiting the applicability of these findings to in vivo mechanisms.
62
Nutrients 2019, 11, 364
rich in facultative anaerobes such as E. Coli, but the faecal microbiota becomes more diverse with the
appearance of obligate anaerobes such as Bifidobacterium and Clostridium within the first week [104].
In a cohort of 19 healthy breastfed full-term Japanese infants, the averaged percentage of obligate
anaerobic bacteria in the gut progressed from 32% (1 day), 37% (7 days), 54% (1 month), 70% (3 months),
64% (6 months), to 99% at 3 years of age. Significant individual variations within this homogenous
cohort diminished by 3 years of age [105,106]. This study did not specify the delivery modes of
this cohort, and the consequent possibility of significant differences in the colonisation patterns of
facultative and obligate anaerobes.
The effects of breastmilk and formula feeding on the infant microbiome and immunity are a
popular topic of research. Breastfeeding has been associated with a decreased risk of necrotising
enterocolitis, infections, and diarrhoea in early life and with a lower incidence of inflammatory bowel
disease, type 2 diabetes, and obesity later in life compared to formula-fed infants [107]. Another
meta-analysis found no association between breastmilk consumption and allergy, asthma, high blood
pressure, or high cholesterol [108]. Considering the complexity of the immune-modulating factors
of breastmilk, the identifying characteristics of the microbiome that contribute to these benefits is
challenging. Bifidobacterium has consistently been found to exist in higher abundances in exclusively
breastfed infants, whereas Lactobacillus has been reported to be higher in formula-fed infants in some
studies [102,109], while at other times reported to be higher in breastfed infants [110]. Backhed et al.
associated exclusive breastfeeding with lower phylogenetic diversity dominated by Bifidobacterium
and Lactobacillus and lower relative abundances of Clostridiales and Bacteroides compared to mixed-fed
infants [7]. Some of these differences may persist throughout the weaning phase as breastmilk and
formula feeding continue with supplementation of solid foods.
In an effort to impart similar bifidogenic effects on formula-fed infants, the supplementation
of infant formula with prebiotics, or prebiotics and probiotics, has become common. A 9:1 ratio
of synthetic linear polymers of GOS:FOS is standard, but these prebiotics represent a simplistic
uniform version of the HMO structures found in breastmilk [20]. Abrahamse-Berkeveld et al. (2016)
found that a combination of short chain GOS (scGOS dp of 3–15), long chain FOS (lcFOS dp of 3–6),
and Bifidobacterium breve increased the abundance of Bifidobacterium from 48% to 60% of the total
bacterial species and reduced the percentage of Clostridium lituseburense/C. histolyticum from 2.6% to
2.0%. [46]. In an in vitro study, Leder et al. (1999) found that many different strains of Bifidobacterium
are capable of utilizing scGOS, but of the species analysed, only B. adolescentis can utilise lcFOS,
providing evidence of the selectivity between related commensal strains and prebiotic structures [111].
These investigations into the utilisation of HMOs and prebiotics in formula offer a starting point for
exploring the effects of prebiotics provided by whole complementary foods.
Oligosaccharides also provide additional protection against pathogenic infection by acting as
structural mimics of the pathogen binding sites that coat the surface of intestinal epithelial cells.
Pathogenic bacteria such as E. Coli bind to oligosaccharides in the lumen, reducing the pathogen load
available for adhesion to intestinal epithelial cells. In Caco-2 and human epithelial type 2 (Hep-2) cell
lines, purified GOS reduced adhesion by 70% and 65% respectively. This effect was dose-dependent
and reached a maximum at 16 mg/mL [45]. It is unclear if complex starches, such as resistant starch,
have the same effect.
63
Nutrients 2019, 11, 364
The community structure and metabolic functions of the infant gut microbiota are strongly
influenced by dietary prebiotics. The bifidogenic nature of breastmilk is well-established and has
been attributed to HMOs [112]. HMO consumption has only been identified in select Bacteroides
(Bifidobacterium) and Lactobacillus species, and different species and subspecies have been found to
utilise different protein-substrate binding and enzymatic mechanisms to metabolise HMOs [113,114].
B. longum subsp. infantis, which is enriched in breastfed infants, express an overabundance
of proteins that transport HMO substrates into the cell, where they are broken up into their
constituent sugars before being metabolised. This limits the sugars that are available to other species
within the microbiota [115]. B. bifidum, however, relies on a set of diverse membrane-associated
extracellular glycosyl hydrolases, lacto-N-biosidase and endo-N-acetylgalactosaminidase [116], which
have comparable enzymatic affinities for HMOs but may release monosaccharides such as lactose,
fucose, and sialic acid into the lumen, which become available to other microbes [117].
Glycosylation patterns on HMOs influence the enzymatic activity that microbes employ. B. breve
has been found to have a preference for sialylated HMOs over neutral HMOs, engaging enzymes that
convert HMOs into multiple intracellular products, but it does not internalise the whole molecule [118].
B. longum has numerous genes for carbohydrate utilisation, including 30 glycosyl hydrolases that
are likely involved in HMO degradation, though adult strains have indicated a preference for plant
polysaccharides [119]. The transcriptomic analysis of B. longum SC596 when shifting from a neutral
HMO substrate to a fucosylated HMO substrate found the gene expression was altered to resemble
the intracellular import strategy of B. infantis, which may provide an example of the facultative gene
expression of infant microbiota in response to dietary factors [20]. A meta-transcriptomic analysis of
faecal samples from a single breastfed baby followed from birth to six months of age, during which
formula, dairy, and solid foods were introduced, found that the carbohydrate fermentation activity of
Bifidobacterium, based on β-galactosidase activity, decreases during weaning while that of the resident
Firmicutes increases, which corresponds with changes in relative abundance of major and minor
species [120].
At approximately 3 months of age, genes implicated in complex carbohydrate utilization are
enriched compared to meconium samples, which favour lactose/galactose and sucrose uptake and
utilisation based on a metagenomic analysis [6]. Just prior to introducing solid foods between 4
and 6 months of age, the gut microbiome derives energy through the degradation of simple sugars,
lactose-specific transport, and carbohydrate uptake, as is expected for a milk-based diet. However,
functional genes involved in plant-polysaccharide metabolism are present prior to the introduction
of complementary weaning foods [6]. By 12 months of age, the infant microbiome is highly enriched
with species and genes active in the degradation of complex sugars and starches [7]. For instance,
Bacteroides thetaiotaomicron, an anaerobic glycan degrading enzyme producer of the Bacteroidetes
phylum, can typically be detected at 12 months of age [6]. An additional study showed that the
increased abundance of Bifidobacterium and decreased abundance of Bacteroides and Clostridium in
breastfed infants compared to formula-fed and mixed-fed infants persists into the weaning phase [121].
Thompson et al. identified differences before and after the introduction of solid foods between
the microbiomes of exclusively breastfed and non-exclusively breastfed infants [122]. Veillonella,
Roseburia, and members of the Lachnospiraceae family appeared with the introduction of solids in
breastfed infants, whereas Streptococcus and Coprobacillus were identified after the introduction of
solids in non-exclusively breastfed infants [122]. Most notable of these findings was the increased
relative abundance of Bifidobacterium after the inclusion of solids in non-exclusively breastfed infants,
compared to a decreased relative abundance of Bifidobacterium after the inclusion of solids in exclusively
breastfed infants, which may reflect differential effects of dietary oligosaccharides and starches during
complementary feeding. Metabolic inferences using a PiCRUST analysis of this limited 16S dataset
showed that only 24 gene clusters encoding enzymes were overrepresented in exclusively breastfed
infants after the introduction of solid foods, including polysaccharide degradation, compared to 230
enzymatic gene clusters overrepresented in the non-exclusively breastfed microbiome, which were
64
Nutrients 2019, 11, 364
primarily involved in signal transduction regulatory systems [122]. This finding indicates differences
in metabolic plasticity between exclusively breastfed and non-exclusively breastfed infants, though
it is possible that the substantial immune factors in breastmilk have a stronger effect on which gene
clusters are overrepresented in this small cohort.
Human faecal microbiota may develop the capacity to degrade a specific type of starch (Type III
Resistant Starch) at weaning, as demonstrated in an in vitro fermentation study using faecal inoculum
collected from breastfed and formula–fed infants before and during weaning [123]. However, species
with the potential capacity to degrade starch have been found to be present at birth [6]. From a
metagenome perspective, microbial networks of infants at 4 months are drastically different to those at
12 months, but polysaccharide degradation has been found to be more pronounced after the cessation
of breastfeeding, rather than during the introduction of solid foods in breastfed infants [7]. It is
possible that the cessation of HMO substrates decreases the need for the expression of HMO-degrading
genes and reduces the competitive advantage of species selective for HMOs, allowing species with a
preference for polysaccharide substrates to assume a greater ecological niche. However, neither the
in vitro experiment nor the metagenomic analysis consider the nutrient availability and degradation
that occurs in the proximal regions of the large intestine prior to analysis of the faecal microbiome.
Starch degradation in the large intestine is a cooperative process that includes enzymatic starch
degradation into glucose, glycolysis leading to SCFAs and organic acids, and hydrogen production.
Starch binding capacity and enzyme specificity underpin the ability of amylolytic microbes to
metabolise starch structures [124]. The presence and function of cellulosomes, amylosomes, and starch
utilisation system gene clusters have been investigated in keystone species belonging to the Firmicutes
and Bacteroidetes families. Three broad classes of amylases have been identified in amylolytic bacteria
that hydrolyse starch into D-glucose: α-amylase for α-1,4 linkages, type 1 pullalanase for α-1,6
linkages, and amylopullalanases for α-1,4 and α-1,6 linkages [125]. Stable Isotope Probing (RNA-SIP),
which allows for the tracking of 13 C-isotope labelled carbon utilisation through metabolite production,
has identified complex trophic structures that implicate primary starch degraders, such as Ruminococcus
bromii, in downstream carbon utilization by microbes found in the infant gut such as Prevotella,
Bifidobacterium, and Eubacterium rectale [126]. The association of amylolytic enzymes with the cell wall
and the ability to stabilize large molecules for cleavage may indicate the function of a given microbe
within the trophic network [127,128]. For instance, extracellular protein complexes on Bacteroides
thetaiotamicron imports starch molecules for internal degradation, limiting the extracellular release of
mono and di-saccharides, compared to outer membrane protein complexes on Clostridium butyricum
which degrade starches outside of the cell before importing the mono- and disaccharides for subsequent
metabolism into SCFAs [47,129,130]. This variety of enzyme structures and systems points to the
metabolic flexibility, which may be increased during dietary transitions such as weaning, that the
microbiome utilises to maximise energy harvest.
Fermentation profiles vary by substrate structure, which changes throughout enzymatic
degradation. Short oligosaccharide chains, such as scFOS, are more rapidly fermented than long
oligosaccharide chains, such as inulin [131]. The rate of fermentation as measured by the SCFA
production was highest during the first 4 hours in a faecal inoculum provided with scFOS substrate,
whereas long chain inulin produced the most SCFA between 12–24 h [131]. Warren et al. (2018)
expanded upon these findings by comparing digested to non-digested starches from a range of
processed, un-processed, digested, and un-digested starch and resistant starch substrates. This study
found that microbiota are able to ferment amorphous and crystalline starches equally well, perhaps
attributable to the range of amylolytic enzymes found in the microbiome, and found no difference
in the fermentation rates of the digested versus undigested substrates [132]. Both the 16S rRNA
gene amplicon sequencing analysis of the inoculum and the SCFA analysis revealed differentiations
according to time-points depending upon the classification of the starch substrate [132].
65
Nutrients 2019, 11, 364
4.4. SCFAs
SCFAs are the primary class of microbial metabolites of starch degradation and are implicated in
immune regulation. SCFAs function as an energy source for the host epithelium and other microbes,
affect lipid metabolism, protect against infection, have anti-inflammatory properties, influence the
gut-brain axis, facilitate immune cell metabolic reprogramming, and regulate immune cell transcription
through epigenetic modifications [133]. SCFA production varies throughout the colon because of
substrate availability, population dynamics, and microbial cross-feeding [134]. The fermentation of
starch substrates by the gut microbial community is characterised by high acetate production, followed
by propionate and relatively less butyrate, though ratios are highly variable [132,135]. RNA-SIP
studies show that lactate is a precursor to both acetate and propionate and that acetate is precursor for
butyrate via both the Co-A transferase pathway and the butyrate kinase pathway [136]. For example,
Bifidobacterium adolescentis can degrade resistant starch leading to the byproducts lactate and acetate.
Actetate is, in turn, used by Eubacterium spp., Roseburia spp., and Coprococcus catus, resulting in the
production of butyrate. Faecalibacterium prausnitzii, an abundant butyrate producer in adults, has not
been detected in infants younger than approximately 2 years of age [137]. Figure 4 demonstrates a
simplified ecological network in which multiple species of bacteria commonly identified in infants
perform parts of the metabolic pathways leading to biosynthesis of SCFAs.
Figure 4. A simplified schematic of the biosynthesis of SCFAs by microbial species identified in human
infants. Organic acid metabolites are outlined, and SCFAs are highlighted in black boxes. Species
of bacteria found in the infant gut microbiome that are implicated in the corresponding pathway
are italicised.
SCFAs begin shaping the enteric environment with the introduction of breastmilk and formula.
Exclusive breastfeeding is associated with lower absolute concentrations of all SCFAs, except
lactate [105]. Ratios of SCFAs within total concentrations have been found to be variable: exclusively
breastfed infants are more likely to have higher proportions of acetate, while partially breastfed and
formula-fed infants are more likely to have relatively higher proportions of propionate, and exclusively
formula-fed infants are likely to have relatively higher proportions of butyrate [138]. However,
measuring SCFAs in faecal samples only provides an indicator of the balance between SCFA production
and absorption. Absorption is likely to vary with epithelial barrier integrity and maturity, which is
known to be influenced by other factors in breastmilk [58,139].
SCFAs modulate immune factors through multiple mechanisms. They increase the expression
of antimicrobial peptides excreted by epithelial cells; modulate the production of cytokines and
66
Nutrients 2019, 11, 364
chemokines; regulate the differentiation, recruitment, and activation of immune cells; and modulate
the differentiation of T lymphocytes [21]. Commonly cited anti-inflammatory properties of SCFAs
can be attributed to their ability to reduce the production and activity of pro-inflammatory cytokines
such as TNF-α and IL-12, often by modulating activity of neutrophils, DCs, and macrophages [140].
Alternatively, SCFAs increase the production of other cytokines, such as IL-18, which has been
implicated in the repair and maintenance of epithelial integrity [141].
Acetate is a minor energy source for gut epithelial cells, a major energy source for muscles
and brain tissue, has anti-inflammatory properties, decreases the pH of the colon, and is used
by cross-feeding species as a co-substrate to produce butyrate [139,142]. Numerous species of
Bifidobacterium readily produce acetate from starchy substrates. Anti-inflammatory properties of acetate
have been linked to the SCFA-dependent modulation of NF-κB in the COLO320DM adenocarcinoma
cell lines, to decreased IL-6 protein release from organ culture, and to decreased LPS-stimulated
TNFα from neutrophils. However, these dose-dependent effects are less pronounced for acetate than
propionate and butyrate [143]. Acetate has also been identified as an important metabolite by which
some subspecies of Bifidobacterium protect against infection, possibly by inhibiting the translocation of
toxins from the gut lumen to the bloodstream [144]. Several in vitro studies suggest that the benefits of
acetate are largely due to the enhanced epithelial integrity, which imparts protection from infection and
inflammation. For instance, B. longum infantis 157F, which is found in breastfed infants and metabolises
glucose to acetate, was found to protect against harmful protein translocation across a Caco-2 epithelial
barrier in an in vitro cell-culture experiment [144].
Propionate has been associated with health benefits most particularly in adults [145]. Similar to
acetate, propionate is a minor energy source for gut epithelial cells, decreases the pH of the colon,
is anti-inflammatory, and has immune modulating properties that in vitro studies of TER in Caco-2
cell lines suggest are linked to beneficial effects on epithelial barrier integrity [146]. Additionally,
propionate decreases liver lipogenesis, serum cholesterol levels, and colorectal carcinogenesis in other
tissues. Insulin sensitivity improvements and increased satiety in adults has also been correlated with
increased propionate levels [139,142,145]. These effects have not been investigated in weaning infants.
Butyrate is the preferred energy source for gut epithelial cells, meaning that little butyrate
reaches systemic circulation. Butyrate also decreases the pH of the colon, promotes epithelial
proliferation, prevents colorectal cancer cell proliferation, reduces oxidative stress, is anti-inflammatory,
and improves gut barrier function by stimulating the production of mucins, antimicrobial peptides,
and TJ proteins [139,142]. Gantois et al. found that butyrate also downregulates the expression
of virulence genes in Salmonella enterica and typhimurium [147]. Butyrate producing bacteria, such
as Eubacterium rectale, Roseburia spp, and Faecalibacterium prausnitzii frequently utilise acetate as a
substrate [148]. The effects of butyrate have been found to be paradoxical where low concentrations
of butyrate (2 mM) promote gut barrier function, characterised by increased TER and decreased
mannitol flux, but high doses (8 mM) may induce cell apoptosis and disrupt the intestinal barrier,
as is characteristic of necrotising enteric colitis [149]. One study identified the benefits of butyrate
to be characteristic of cellular differentiation because of the increased dome formation and alkaline
phosphatase activity [146], whereas another identified cell migration, as is needed for epithelial repair,
as a beneficial mechanism [149]. Both studies found that the effects were dependent on protein
synthesis and gene transcription but not the beta-oxidation or activation of adenosine 3’, 5’-cyclic
monophosphate [146,149].
Most investigations into SCFAs have focused on adult populations. In infants, SCFAs are
considered beneficial, but faecal measurements are inappropriate to use as an indicator of a healthy
microbiome due to its paradoxical effects at high concentrations and the importance of considering the
balance of SCFA utilisation by epithelial cells and absorption into the blood stream.
67
Nutrients 2019, 11, 364
4.5. Vitamins
Vitamins are an additional class of secondary metabolites produced by the microbiota with effects
on immunity. Commensal bacteria have the capacity to synthesise essential vitamins, particularly from
the B and K groups, the expression of which is distinct in infants compared to adults. The microbiota in
neonates demonstrate the enrichment of genes involved in the production of Vitamin K2, retinol, folate,
pyroxidal (B6), and biotin (B7), which are involved in bone, vision, tooth development, and glucose
conversion are upregulated in the neonatal microbiome. Genes involved in the transport of B12, iron,
hemin, and heme are also enriched in neonates but decline markedly with age, corresponding with
increased nutritional demands for iron between 4 and 6 months of age. Throughout the weaning
months, genes involved in the biosynthesis of thiamine (B1), pantothenate (B5), cobalamin (B12),
and lysine increase [150]. Methionine degradation and leucine and tryptophan biosynthesis increase
to reach levels comparable to mothers by 12 months of age [7].
It has been estimated that B vitamins are produced by 40–65% of human gut microbes, with
some microbes having the capacity to produce all 8 B-vitamins and some demonstrating pathways
that complemented those of other organisms [151]. These estimates were determined by aligning
metagenomes from the human gut microbiome to an annotated genome on the PubSEED platform [151].
Bifidobacteriales contained the most conserved pattern of B1, B3, and B7 in approximately 35% of the
genomes, whereas Bacteroidetes demonstrated biosynthetic pathways of all 8 vitamins present in
51% of the genomes [151]. Prevotellaceaes produce B2, B5, and B7; Lactobacillales either contain no
biosynthetic pathways, or were limited to B2; and Clostridiales produce only B12 [151]. The full folate
biosynthesis (B9) pathway is present in 43% of genomes, which is distributed in nearly all Bacteroidetes
genomes, in most Fusobacteria and Proteobacteria, and in partial pathways occuring in Actinobacteria and
Firmicutes [151]. Vitamin K in one of two forms is reported to be produced by Bacteroides, Enterobacter,
Veillonella, and Eubacterium lentum, though the bioavailability of bacterially-derived Vitamin K has not
been established [152]. How these genes are differentially expressed in the infant microbiome and in
response to specific types of complementary foods has yet to be explored.
The interactions between microbially-derived vitamins and immune cells are varied and poorly
characterised. Of the known pathways, B6 has been found to serve as a cofactor in immunomodulatory
pathways [153], B9 has been implicated in the maintenance of regulatory T cells [154], and B12 has been
found to augment CD8+ T lymphocytes and NK cell activity in deficient patients [35,155]. Interestingly,
the byproducts from vitamin synthesis pathways have also recently been implicated in immune cell
recognition: mucosa-associated invariant T cells, which produce IL-17 and IFN-γ, are activated in
response to microbe-derived products of the riboflavin biosynthetic pathway that are presented by
a monomeric major histocompatibility complex class 1 (MHC-1)-like related molecule (MR1) [156].
These MHC and MHC-like molecules are imperative to discriminate self from non-self, enabling
protective immunity [157].
68
Nutrients 2019, 11, 364
Author Contributions: Conceptualisation, S.M., W.Y., and N.C.R.; resources, N.C.R. and W.C.M.;
writing—original draft preparation, S.M.; writing—review and editing, W.Y., K.F., J.M., N.C.R., and W.C.M.;
supervision, W.Y., K.F., W.C.M., and N.C.R.; project administration, N.C.R. and W.C.M.; funding acquisition,
N.C.R.
Funding: This research received no external funding.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Sen, P.; Mardinogulu, A.; Nielsen, J. Selection of complementary foods based on optimal nutritional values.
Sci. Rep. 2017, 7, 5413. [CrossRef] [PubMed]
2. Lonnerdal, B.; Hernell, O. An opinion on “staging” of infant formula: A developmental perspective on infant
feeding. J. Pediatr. Gastroenterol. Nutr. 2016, 62, 9–21. [CrossRef] [PubMed]
3. Krebs, N.F.; Hambidge, K.M. Zinc requirements and zinc intakes of breast-fed infants. Am. J. Clin. Nutr.
1986, 43, 288–292. [CrossRef] [PubMed]
4. Young, B.E.; Krebs, N.F. Complementary feeding: Critical considerations to optimize growth, nutrition,
and feeding behavior. Curr. Pediatr. Rep. 2013, 1, 247–256. [CrossRef] [PubMed]
5. Cong, X.; Xu, W.; Janton, S.; Henderson, W.A.; Matson, A.; McGrath, J.M.; Maas, K.; Graf, J. Gut microbiome
developmental patterns in early life of preterm infants: Impacts of feeding and gender. PloS ONE 2016, 11,
e0152751. [CrossRef] [PubMed]
6. Koenig, J.E.; Spor, A.; Scalfone, N.; Fricker, A.D.; Stombaugh, J.; Knight, R.; Angenent, L.T.; Ley, R.E.
Succession of microbial consortia in the developing infant gut microbiome. Proc. Natl. Acad. Sci. USA 2011,
108 (Suppl. 1), 4578–4585. [CrossRef]
7. Backhed, F.; Roswall, J.; Peng, Y.; Feng, Q.; Jia, H.; Kovatcheva-Datchary, P.; Li, Y.; Xia, Y.; Xie, H.; Zhong, H.;
et al. Dynamics and stabilization of the human gut microbiome during the first year of life. Cell Host Microbe
2015, 17, 852. [CrossRef]
8. Hill, C.J.; Lynch, D.B.; Murphy, K.; Ulaszewska, M.; Jeffery, I.B.; O’Shea, C.A.; Watkins, C.; Dempsey, E.;
Mattivi, F.; Tuohy, K. Evolution of gut microbiota composition from birth to 24 weeks in the infantmet cohort.
Microbiome 2017, 5, 4. [CrossRef]
9. Goulet, O. Potential role of the intestinal microbiota in programming health and disease. Nutr. Rev. 2015, 73
(Suppl. 1), 32–40. [CrossRef]
10. Praveen, P.; Jordan, F.; Priami, C.; Morine, M.J. The role of breast-feeding in infant immune system: A systems
perspective on the intestinal microbiome. Microbiome 2015, 3, 41. [CrossRef] [PubMed]
11. Walker, W.A. Initial intestinal colonization in the human infant and immune homeostasis. Ann. Nutr. Metab.
2013, 63 (Suppl. 2), 8–15. [CrossRef]
12. Amenyogbe, N.; Kollmann, T.R.; Ben-Othman, R. Early-life host-microbiome interphase: The key frontier for
immune development. Front. Pediatr. 2017, 5, 111. [CrossRef]
69
Nutrients 2019, 11, 364
13. Clavel, T.; Gomes-Neto, J.C.; Lagkouvardos, I.; Ramer-Tait, A.E. Deciphering interactions between the gut
microbiota and the immune system via microbial cultivation and minimal microbiomes. Immunol. Rev. 2017,
279, 8–22. [CrossRef] [PubMed]
14. Kaplan, J.L.; Shi, H.N.; Walker, W.A. The role of microbes in developmental immunologic programming.
Pediatr. Res. 2011, 69, 465–472. [CrossRef]
15. Li, M.; Wang, M.; Donovan, S.M. Early development of the gut microbiome and immune-mediated childhood
disorders. Semin. Reprod. Med. 2014, 32, 74–86. [CrossRef] [PubMed]
16. Bertelsen, R.J.; Jensen, E.T.; Ringel-Kulka, T. Use of probiotics and prebiotics in infant feeding. Best Pract. Res.
Clin. Gastroenterol. 2016, 30, 39–48. [CrossRef]
17. Blacher, E.; Levy, M.; Tatirovsky, E.; Elinav, E. Microbiome-modulated metabolites at the interface of host
immunity. J. Immunol. 2017, 198, 572–580. [CrossRef] [PubMed]
18. Kwak, M.J.; Kwon, S.K.; Yoon, J.K.; Song, J.Y.; Seo, J.G.; Chung, M.J.; Kim, J.F. Evolutionary architecture
of the infant-adapted group of bifidobacterium species associated with the probiotic function. Syst. Appl.
Microbiol. 2016, 39, 429–439. [CrossRef]
19. Sierra, C.; Bernal, M.J.; Blasco, J.; Martinez, R.; Dalmau, J.; Ortuno, I.; Espin, B.; Vasallo, M.I.; Gil, D.;
Vidal, M.L.; et al. Prebiotic effect during the first year of life in healthy infants fed formula containing gos as
the only prebiotic: A multicentre, randomised, double-blind and placebo-controlled trial. Eur. J. Nutr. 2015,
54, 89–99. [CrossRef]
20. Thomson, P.; Medina, D.A.; Garrido, D. Human milk oligosaccharides and infant gut bifidobacteria:
Molecular strategies for their utilization. Food Microbiol. 2018, 75, 37–46. [CrossRef] [PubMed]
21. Corrêa-Oliveira, R.; Fachi, J.L.; Vieira, A.; Sato, F.T.; Vinolo, M.A.R. Regulation of immune cell function by
short-chain fatty acids. Clin. Transl. Immunol. 2016, 5, e73. [CrossRef] [PubMed]
22. Maier, T.V.; Lucio, M.; Lee, L.H.; VerBerkmoes, N.C.; Brislawn, C.J.; Bernhardt, J.; Lamendella, R.;
McDermott, J.E.; Bergeron, N.; Heinzmann, S.S.; et al. Impact of dietary resistant starch on the human gut
microbiome, metaproteome, and metabolome. mBio 2017, 8. [CrossRef] [PubMed]
23. Tanaka, M.; Nakayama, J. Development of the gut microbiota in infancy and its impact on health in later life.
Allergol. Int. 2017, 66, 515–522. [CrossRef] [PubMed]
24. Schroeder, B.O.; Birchenough, G.M.; Ståhlman, M.; Arike, L.; Johansson, M.E.; Hansson, G.C.; Bäckhed, F.
Bifidobacteria or fiber protects against diet-induced microbiota-mediated colonic mucus deterioration.
Cell Host Microbe 2018, 23, 27–40. [CrossRef] [PubMed]
25. Rogier, E.W.; Frantz, A.L.; Bruno, M.E.; Wedlund, L.; Cohen, D.A.; Stromberg, A.J.; Kaetzel, C.S. Lessons
from mother: Long-term impact of antibodies in breast milk on the gut microbiota and intestinal immune
system of breastfed offspring. Gut Microbes 2014, 5, 663–668. [CrossRef] [PubMed]
26. Rogier, E.W.; Frantz, A.L.; Bruno, M.E.; Wedlund, L.; Cohen, D.A.; Stromberg, A.J.; Kaetzel, C.S. Secretory
antibodies in breast milk promote long-term intestinal homeostasis by regulating the gut microbiota and
host gene expression. Proc. Natl. Acad. Sci. USA 2014, 111, 3074–3079. [CrossRef] [PubMed]
27. Cui, X.; Li, Y.; Yang, L.; You, L.; Wang, X.; Shi, C.; Ji, C.; Guo, X. Peptidome analysis of human milk from
women delivering macrosomic fetuses reveals multiple means of protection for infants. Oncotarget 2016, 7,
63514–63525. [CrossRef]
28. Duranti, S.; Lugli, G.A.; Mancabelli, L.; Armanini, F.; Turroni, F.; James, K.; Ferretti, P.; Gorfer, V.; Ferrario, C.;
Milani, C.; et al. Maternal inheritance of bifidobacterial communities and bifidophages in infants through
vertical transmission. Microbiome 2017, 5, 66. [CrossRef]
29. Gonzalez-Perez, G.; Hicks, A.L.; Tekieli, T.M.; Radens, C.M.; Williams, B.L.; Lamouse-Smith, E.S. Maternal
antibiotic treatment impacts development of the neonatal intestinal microbiome and antiviral immunity. J.
Immunol. 2016, 196, 3768–3779. [CrossRef]
30. Hartwig, I.; Diemert, A.; Tolosa, E.; Hecher, K.; Arck, P. Babies galore; or recent findings and future
perspectives of pregnancy cohorts with a focus on immunity. J. Reprod. Immunol. 2015, 108, 6–11. [CrossRef]
31. Romano-Keeler, J.; Weitkamp, J.H. Maternal influences on fetal microbial colonization and immune
development. Pediatri. Res. 2015, 77, 189–195. [CrossRef] [PubMed]
32. Schlinzig, T.; Johansson, S.; Stephansson, O.; Hammarstrom, L.; Zetterstrom, R.H.; von Dobeln, U.;
Cnattingius, S.; Norman, M. Surge of immune cell formation at birth differs by mode of delivery and
infant characteristics-a population-based cohort study. PloS ONE 2017, 12, e0184748. [CrossRef]
70
Nutrients 2019, 11, 364
33. Schwartz, C.; Chabanet, C.; Szleper, E.; Feyen, V.; Issanchou, S.; Nicklaus, S. Infant acceptance of primary
tastes and fat emulsion: Developmental changes and links with maternal and infant characteristics.
Chem. Senses 2017, 42, 593–603. [CrossRef] [PubMed]
34. Belkaid, Y.; Segre, J.A. Dialogue between skin microbiota and immunity. Science 2014, 346, 954–959.
[CrossRef] [PubMed]
35. Brestoff, J.R.; Artis, D. Commensal bacteria at the interface of host metabolism and the immune system.
Nat. Immunol. 2013, 14, 676–684. [CrossRef] [PubMed]
36. Kabat, A.M.; Srinivasan, N.; Maloy, K.J. Modulation of immune development and function by intestinal
microbiota. Trends Immunol. 2014, 35, 507–517. [CrossRef]
37. Martin, R.; Nauta, A.J.; Ben Amor, K.; Knippels, L.M.; Knol, J.; Garssen, J. Early life: Gut microbiota and
immune development in infancy. Benef. Microbes 2010, 1, 367–382. [CrossRef]
38. Paun, A.; Danska, J.S. Immuno-ecology: How the microbiome regulates tolerance and autoimmunity.
Curr. Opin. Immunol. 2015, 37, 34–39. [CrossRef]
39. Ruff, W.E.; Kriegel, M.A. Autoimmune host-microbiota interactions at barrier sites and beyond. Trends Mol.
Med. 2015, 21, 233–244. [CrossRef]
40. Shi, N.; Li, N.; Duan, X.; Niu, H. Interaction between the gut microbiome and mucosal immune system.
Mil. Med. Res. 2017, 4, 14. [CrossRef]
41. Sjogren, Y.M.; Tomicic, S.; Lundberg, A.; Bottcher, M.F.; Bjorksten, B.; Sverremark-Ekstrom, E.; Jenmalm, M.C.
Influence of early gut microbiota on the maturation of childhood mucosal and systemic immune responses.
Clin. Exp. Allergy 2009, 39, 1842–1851. [CrossRef] [PubMed]
42. Turroni, F.; Milani, C.; Duranti, S.; Mancabelli, L.; Mangifesta, M.; Viappiani, A.; Lugli, G.A.; Ferrario, C.;
Gioiosa, L.; Ferrarini, A.; et al. Deciphering bifidobacterial-mediated metabolic interactions and their impact
on gut microbiota by a multi-omics approach. ISME J. 2016, 10, 1656–1668. [CrossRef] [PubMed]
43. Triantis, V.; Bode, L.; Van Neerven, R. Immunological effects of human milk oligosaccharides. Front. Pediatr.
2018, 6, 190. [CrossRef] [PubMed]
44. Prell, C.; Koletzko, B. Breastfeeding and complementary feeding: Recommendations on infant nutrition.
Deutsches Ärzteblatt Int. 2016, 113, 435. [PubMed]
45. Shoaf, K.; Mulvey, G.L.; Armstrong, G.D.; Hutkins, R.W. Prebiotic galactooligosaccharides reduce adherence
of enteropathogenic escherichia coli to tissue culture cells. Infect. Immun. 2006, 74, 6920–6928. [CrossRef]
[PubMed]
46. Abrahamse-Berkeveld, M.; Alles, M.; Franke-Beckmann, E.; Helm, K.; Knecht, R.; Kollges, R.; Sandner, B.;
Knol, J.; Ben Amor, K.; Bufe, A. Infant formula containing galacto-and fructo-oligosaccharides and
bifidobacterium breve m-16v supports adequate growth and tolerance in healthy infants in a randomised,
controlled, double-blind, prospective, multicentre study. J. Nutr. Sci. 2016, 5, e42. [CrossRef]
47. Swennen, K.; Courtin, C.M.; Delcour, J.A. Non-digestible oligosaccharides with prebiotic properties. Crit. Rev.
Food Sci. Nutr. 2006, 46, 459–471. [CrossRef]
48. Dallman, P. Nutritional anemia of infancy: Iron, folic acid, and vitamin B12. Nutri. Infancy Phila. Henley
Belfus Inc 1988, 216–235.
49. Zealand, N. Food and Nutrition Guidelines for Healthy Infants and Toddlers (Aged 0-2): Background Paper; Ministry
of Health: Wellington, New Zealand, 2008.
50. Allen, L.H. B vitamins in breast milk: Relative importance of maternal status and intake, and effects on
infant status and function. Adv. Nutr. 2012, 3, 362–369. [CrossRef]
51. Schack-Nielsen, L.; Sorensen, T.; Mortensen, E.L.; Michaelsen, K.F. Late introduction of complementary
feeding, rather than duration of breastfeeding, may protect against adult overweight. Am. J. Clin. Nutr. 2010,
91, 619–627. [CrossRef]
52. Ong, K.K.; Kennedy, K.; Castaneda-Gutierrez, E.; Forsyth, S.; Godfrey, K.M.; Koletzko, B.; Latulippe, M.E.;
Ozanne, S.E.; Rueda, R.; Schoemaker, M.H.; et al. Postnatal growth in preterm infants and later health
outcomes: A systematic review. Acta Paediatr. 2015, 104, 974–986. [CrossRef] [PubMed]
53. Pluymen, L.P.; Wijga, A.H.; Gehring, U.; Koppelman, G.H.; Smit, H.A.; van Rossem, L. Early introduction of
complementary foods and childhood overweight in breastfed and formula-fed infants in the netherlands:
The piama birth cohort study. Eur. J. Nutr. 2018, 1–9. [CrossRef] [PubMed]
54. Georgi, G.; Bartke, N.; Wiens, F.; Stahl, B. Functional glycans and glycoconjugates in human milk. Am. J.
Clin. Nutr. 2013, 98, 578S–585S. [CrossRef] [PubMed]
71
Nutrients 2019, 11, 364
55. Engfer, M.B.; Stahl, B.; Finke, B.; Sawatzki, G.; Daniel, H. Human milk oligosaccharides are resistant to
enzymatic hydrolysis in the upper gastrointestinal tract. Am. J. Clin. Nutr. 2000, 71, 1589–1596. [CrossRef]
[PubMed]
56. Oozeer, R.; van Limpt, K.; Ludwig, T.; Ben Amor, K.; Martin, R.; Wind, R.D.; Boehm, G.; Knol, J. Intestinal
microbiology in early life: Specific prebiotics can have similar functionalities as human-milk oligosaccharides.
Am. J. Clin. Nutr. 2013, 98, 561S–571S. [CrossRef] [PubMed]
57. Boehm, G.; Moro, G. Structural and functional aspects of prebiotics used in infant nutrition. J. Nutr. 2008,
138, 1818S–1828S. [CrossRef] [PubMed]
58. Cummins, A.; Thompson, F. Effect of breast milk and weaning on epithelial growth of the small intestine in
humans. Gut 2002, 51, 748–754. [CrossRef]
59. Keusch, G.T.; Denno, D.M.; Black, R.E.; Duggan, C.; Guerrant, R.L.; Lavery, J.V.; Nataro, J.P.; Rosenberg, I.H.;
Ryan, E.T.; Tarr, P.I. Environmental enteric dysfunction: Pathogenesis, diagnosis, and clinical consequences.
Clin. Infect. Dis. 2014, 59, S207–S212. [CrossRef]
60. Bourlieu, C.; Menard, O.; Bouzerzour, K.; Mandalari, G.; Macierzanka, A.; Mackie, A.R.; Dupont, D.
Specificity of infant digestive conditions: Some clues for developing relevant in vitro models. Crit. Rev. Food
Sci. Nutr. 2014, 54, 1427–1457. [CrossRef]
61. Sibley, E. Carbohydrate intolerance. Curr. Opin. Gastroenterol. 2004, 20, 162–167. [CrossRef]
62. Lebenthal, E.; Lee, P.C.; Heitlinger, L.A. Impact of development of the gastrointestinal tract on infant feeding.
J. Pediatr. 1983, 102, 1–9. [CrossRef]
63. Lin, A.H.-M.; Nichols, B.L. The digestion of complementary feeding starches in the young child. Starch 2017,
69. [CrossRef]
64. Parracho, H.; McCartney, A.L.; Gibson, G.R. Probiotics and prebiotics in infant nutrition. Proc. Nutr. Soc.
2007, 66, 405–411. [CrossRef] [PubMed]
65. Dewit, O.; Dibba, B.; Prentice, A. Breast-milk amylase activity in english and gambian mothers: Effects of
prolonged lactation, maternal parity, and individual variations. Pediatr. Res. 1990, 28, 502–506. [CrossRef]
[PubMed]
66. Anderson, J.M.; Van Itallie, C.M. Physiology and function of the tight junction. Cold Spring Harb. Perspect.
Biol. 2009, 1, a002584. [CrossRef] [PubMed]
67. Georas, S.N.; Rezaee, F. Epithelial barrier function: At the front line of asthma immunology and allergic
airway inflammation. J. Allergy Clin. Immunol. 2014, 134, 509–520. [CrossRef] [PubMed]
68. Denno, D.M.; VanBuskirk, K.; Nelson, Z.C.; Musser, C.A.; Hay Burgess, D.C.; Tarr, P.I. Use of the lactulose to
mannitol ratio to evaluate childhood environmental enteric dysfunction: A systematic review. Clin. Infect.
Dis. 2014, 59, S213–S219. [CrossRef]
69. Rugtveit, J.; Fagerhol, M.K. Age-dependent variations in fecal calprotectin concentrations in children.
J. Pediatr. Gastroenterol. Nutr. 2002, 34, 323. [CrossRef]
70. Dorosko, S.M.; MacKenzie, T.; Connor, R.I. Fecal calprotectin concentrations are higher in exclusively
breastfed infants compared to those who are mixed-fed. Breastfeed. Med. 2008, 3, 117–119. [CrossRef]
71. De Walle, J.V.; Sergent, T.; Piront, N.; Toussaint, O.; Schneider, Y.-J.; Larondelle, Y. Deoxynivalenol affects
in vitro intestinal epithelial cell barrier integrity through inhibition of protein synthesis. Toxicol. Appl.
Pharmacol. 2010, 245, 291–298. [CrossRef]
72. Akbari, P.; Braber, S.; Alizadeh, A.; Verheijden, K.A.; Schoterman, M.H.; Kraneveld, A.D.; Garssen, J.;
Fink-Gremmels, J. Galacto-oligosaccharides protect the intestinal barrier by maintaining the tight junction
network and modulating the inflammatory responses after a challenge with the mycotoxin deoxynivalenol
in human caco-2 cell monolayers and b6c3f1 mice. J. Nutr. 2015, 145, 1604–1613. [CrossRef] [PubMed]
73. Akbari, P.; Fink-Gremmels, J.; Willems, R.; Difilippo, E.; Schols, H.A.; Schoterman, M.H.C.; Garssen, J.;
Braber, S. Characterizing microbiota-independent effects of oligosaccharides on intestinal epithelial cells:
Insight into the role of structure and size: Structure-activity relationships of non-digestible oligosaccharides.
Eur. J. Nutr. 2017, 56, 1919–1930. [CrossRef] [PubMed]
74. Cornick, S.; Tawiah, A.; Chadee, K. Roles and regulation of the mucus barrier in the gut. Tissue Barriers 2015,
3, e982426. [CrossRef] [PubMed]
75. Turroni, F.; Serafini, F.; Foroni, E.; Duranti, S.; Motherway, M.O.C.; Taverniti, V.; Mangifesta, M.; Milani, C.;
Viappiani, A.; Roversi, T. Role of sortase-dependent pili of bifidobacterium bifidum prl2010 in modulating
bacterium–host interactions. Proc. Natl. Acad. Sci. USA 2013, 110, 11151–11156. [CrossRef] [PubMed]
72
Nutrients 2019, 11, 364
76. Walker, A. Probiotics stick it to the man. Nat. Rev. Microbiol. 2009, 7, 843. [CrossRef] [PubMed]
77. Haange, S.-B.; Oberbach, A.; Schlichting, N.; Hugenholtz, F.; Smidt, H.; von Bergen, M.; Till, H.; Seifert, J.
Metaproteome analysis and molecular genetics of rat intestinal microbiota reveals section and localization
resolved species distribution and enzymatic functionalities. J. Proteome Res. 2012, 11, 5406–5417. [CrossRef]
[PubMed]
78. Zmora, N.; Zilberman-Schapira, G.; Suez, J.; Mor, U.; Dori-Bachash, M.; Bashiardes, S.; Kotler, E.; Zur, M.;
Regev-Lehavi, D.; Brik, R.B.-Z. Personalized gut mucosal colonization resistance to empiric probiotics is
associated with unique host and microbiome features. Cell 2018, 174, 1388–1405.e1321. [CrossRef] [PubMed]
79. Caballero-Franco, C.; Keller, K.; De Simone, C.; Chadee, K. The vsl#3 probiotic formula induces mucin gene
expression and secretion in colonic epithelial cells. Am. J. Physiol. Gastrointest. Liver Physiol. 2007, 292,
G315–G322.
80. Nofrarias, M.; Martinez-Puig, D.; Pujols, J.; Majo, N.; Perez, J.F. Long-term intake of resistant starch improves
colonic mucosal integrity and reduces gut apoptosis and blood immune cells. Nutrition 2007, 23, 861–870.
[CrossRef]
81. Brockhausen, I. Sulphotransferases acting on mucin-type oligosaccharides. Biochem. Soc. Trans. 2003, 31,
318–325. [CrossRef]
82. Shan, M.; Gentile, M.; Yeiser, J.R.; Walland, A.C.; Bornstein, V.U.; Chen, K.; He, B.; Cassis, L.; Bigas, A.;
Cols, M. Mucus enhances gut homeostasis and oral tolerance by delivering immunoregulatory signals.
Science 2013, 1237910. [CrossRef]
83. Yassour, M.; Vatanen, T.; Siljander, H.; Hämäläinen, A.-M.; Härkönen, T.; Ryhänen, S.J.; Franzosa, E.A.;
Vlamakis, H.; Huttenhower, C.; Gevers, D. Natural history of the infant gut microbiome and impact of
antibiotic treatment on bacterial strain diversity and stability. Sci. Transl. Med. 2016, 8, 343ra381. [CrossRef]
[PubMed]
84. Jeurink, P.V.; van Esch, B.C.; Rijnierse, A.; Garssen, J.; Knippels, L.M. Mechanisms underlying immune
effects of dietary oligosaccharides. Am. J. Clin. Nutr. 2013, 98, 572S–577S. [CrossRef] [PubMed]
85. Kollmann, T.R.; Levy, O.; Montgomery, R.R.; Goriely, S. Innate immune function by toll-like receptors:
Distinct responses in newborns and the elderly. Immunity 2012, 37, 771–783. [CrossRef] [PubMed]
86. Goenka, A.; Kollmann, T.R. Development of immunity in early life. J. Infect. 2015, 71 (Suppl. 1), S112–S120.
[CrossRef]
87. Duerkop, B.A.; Vaishnava, S.; Hooper, L.V. Immune responses to the microbiota at the intestinal mucosal
surface. Immunity 2009, 31, 368–376. [CrossRef] [PubMed]
88. Fung, T.C.; Artis, D.; Sonnenberg, G.F. Anatomical localization of commensal bacteria in immune cell
homeostasis and disease. Immunol. Rev. 2014, 260, 35–49. [CrossRef]
89. de Kivit, S.; Tobin, M.C.; Forsyth, C.B.; Keshavarzian, A.; Landay, A.L. Regulation of intestinal immune
responses through tlr activation: Implications for pro- and prebiotics. Front. Immunol. 2014, 5, 60. [CrossRef]
90. Geijtenbeek, T.B.; Gringhuis, S.I. Signalling through c-type lectin receptors: Shaping immune responses. Nat.
Rev. Immunol. 2009, 9, 465. [CrossRef]
91. Macpherson, A.J.; Uhr, T. Induction of protective iga by intestinal dendritic cells carrying commensal bacteria.
Science 2004, 303, 1662–1665. [CrossRef]
92. Belkaid, Y.; Hand, T.W. Role of the microbiota in immunity and inflammation. Cell 2014, 157, 121–141.
[CrossRef] [PubMed]
93. Gonzalez-Perez, G.; Lamouse-Smith, E.S. Gastrointestinal microbiome dysbiosis in infant mice alters
peripheral CD8(+) T cell receptor signaling. Front. Immunol. 2017, 8, 265. [CrossRef] [PubMed]
94. Bermudez-Brito, M.; Rosch, C.; Schols, H.A.; Faas, M.M.; de Vos, P. Resistant starches differentially stimulate
toll-like receptors and attenuate proinflammatory cytokines in dendritic cells by modulation of intestinal
epithelial cells. Mol. Nutr. Food Res. 2015, 59, 1814–1826. [CrossRef] [PubMed]
95. Vogt, L.; Ramasamy, U.; Meyer, D.; Pullens, G.; Venema, K.; Faas, M.M.; Schols, H.A.; de Vos, P. Immune
modulation by different types of beta2–>1-fructans is toll-like receptor dependent. PloS ONE 2013, 8, e68367.
[CrossRef] [PubMed]
96. Bermudez-Brito, M.; Faas, M.M.; de Vos, P. Modulation of dendritic-epithelial cell responses against
sphingomonas paucimobilis by dietary fibers. Sci. Rep. 2016, 6, 30277. [CrossRef] [PubMed]
97. Aagaard, K.; Ma, J.; Antony, K.M.; Ganu, R.; Petrosino, J.; Versalovic, J. The placenta harbors a unique
microbiome. Sci. Transl. Med. 2014, 6, 237ra265. [CrossRef] [PubMed]
73
Nutrients 2019, 11, 364
98. Bager, P.; Wohlfahrt, J.; Westergaard, T. Caesarean delivery and risk of atopy and allergic disesase:
Meta-analyses. Clin. Exp. Allergy 2008, 38, 634–642. [CrossRef] [PubMed]
99. Negele, K.; Heinrich, J.; Borte, M.; von Berg, A.; Schaaf, B.; Lehmann, I.; Wichmann, H.E.; Bolte, G.; LISA
Study Group. Mode of delivery and development of atopic disease during the first 2 years of life. Pediatr.
Allergy Immunol. 2004, 15, 48–54. [CrossRef]
100. Sevelsted, A.; Stokholm, J.; Bønnelykke, K.; Bisgaard, H. Cesarean section and chronic immune disorders.
Pediatrics 2015, 135, e92–e98. [CrossRef] [PubMed]
101. Yuan, C.; Gaskins, A.J.; Blaine, A.I.; Zhang, C.; Gillman, M.W.; Missmer, S.A.; Field, A.E.; Chavarro, J.E.
Association between cesarean birth and risk of obesity in offspring in childhood, adolescence, and early
adulthood. JAMA Pediatr. 2016, 170, e162385. [CrossRef]
102. Penders, J.; Thijs, C.; Vink, C.; Stelma, F.F.; Snijders, B.; Kummeling, I.; van den Brandt, P.A.; Stobberingh, E.E.
Factors influencing the composition of the intestinal microbiota in early infancy. Pediatrics 2006, 118, 511–521.
[CrossRef] [PubMed]
103. Appleman, M.D.; Wormser, G.P. Strict and facultative anaerobes: Medical and environmental aspects edited
by michiko m. Nakano and peter zuber wymondham, norfolk, u.K.: Horizon bioscience, 2004 392 pp.,
illustrated. $139.95 (cloth). Clin. Infect. Dis. 2005, 41, 132. [CrossRef]
104. Jiménez, E.; Marín, M.L.; Martín, R.; Odriozola, J.M.; Olivares, M.; Xaus, J.; Fernández, L.; Rodríguez, J.M. Is
meconium from healthy newborns actually sterile? Res. Microbiol. 2008, 159, 187–193. [CrossRef] [PubMed]
105. Nagpal, R.; Kurakawa, T.; Tsuji, H.; Takahashi, T.; Kawashima, K.; Nagata, S.; Nomoto, K.; Yamashiro, Y.
Evolution of gut bifidobacterium population in healthy japanese infants over the first three years of life:
A quantitative assessment. Sci. Rep. 2017, 7, 10097. [CrossRef] [PubMed]
106. Nagpal, R.; Tsuji, H.; Takahashi, T.; Nomoto, K.; Kawashima, K.; Nagata, S.; Yamashiro, Y. Ontogenesis of the
gut microbiota composition in healthy, full-term, vaginally born and breast-fed infants over the first 3 years
of life: A quantitative bird’s-eye view. Front. Microbiol. 2017, 8, 1388. [CrossRef]
107. Young, B. Breastfeeding and human milk: Short and long-term health benefits to the recipient infant. In Early
Nutrition and Long-Term Health; Saavedra, J.M., Dattilo, A.M., Eds.; Elsevier: Amsterdam, The Netherlands,
2017; pp. 25–53.
108. Victora, C.G.; Bahl, R.; Barros, A.J.; França, G.V.; Horton, S.; Krasevec, J.; Murch, S.; Sankar, M.J.; Walker, N.;
Rollins, N.C. Breastfeeding in the 21st century: Epidemiology, mechanisms, and lifelong effect. Lancet 2016,
387, 475–490. [CrossRef]
109. Penders, J.; Thijs, C.; van den Brandt, P.A.; Kummeling, I.; Snijders, B.; Stelma, F.; Adams, H.; van
Ree, R.; Stobberingh, E.E. Gut microbiota composition and development of atopic manifestations in infancy:
The koala birth cohort study. Gut 2006, 56, 661–667. [CrossRef] [PubMed]
110. Rinne, M.; Kalliomaki, M.; Arvilommi, H.; Salminen, S.; Isolauri, E. Effect of probiotics and breastfeeding on
the bifidobacterium and lactobacillus/enterococcus microbiota and humoral immune responses. J. Pediatr.
2005, 147, 186–191. [CrossRef] [PubMed]
111. Leder, S.; Hartmeier, W.; Marx, S.P. Alpha-galactosidase of bifidobacterium adolescentis dsm 20083. Curr.
Microbiol. 1999, 38, 101–106. [CrossRef] [PubMed]
112. Smilowitz, J.T.; Lebrilla, C.B.; Mills, D.A.; German, J.B.; Freeman, S.L. Breast milk oligosaccharides:
Structure-function relationships in the neonate. Annu. Rev. Nutr. 2014, 34, 143–169. [CrossRef] [PubMed]
113. Marcobal, A.; Sonnenburg, J. Human milk oligosaccharide consumption by intestinal microbiota.
Clin. Microbiol. Infect. 2012, 18, 12–15. [CrossRef] [PubMed]
114. Asakuma, S.; Hatakeyama, E.; Urashima, T.; Yoshida, E.; Katayama, T.; Yamamoto, K.; Kumagai, H.;
Ashida, H.; Hirose, J.; Kitaoka, M. Physiology of the consumption of human milk oligosaccharides by
infant-gut associated bifidobacteria. J. Biol. Chem. 2011, 111, 24583–24592. [CrossRef] [PubMed]
115. Garrido, D.; Kim, J.H.; German, J.B.; Raybould, H.E.; Mills, D.A. Oligosaccharide binding proteins from
bifidobacterium longum subsp. Infantis reveal a preference for host glycans. PloS ONE 2011, 6, e17315.
[CrossRef]
116. Wada, J.; Ando, T.; Kiyohara, M.; Ashida, H.; Kitaoka, M.; Yamaguchi, M.; Kumagai, H.; Katayama, T.;
Yamamoto, K. Bifidobacterium bifidum lacto-n-biosidase, a critical enzyme for the degradation of human
milk oligosaccharides with a type 1 structure. Appl. Environ. Microbiol. 2008, 74, 3996–4004. [CrossRef]
[PubMed]
74
Nutrients 2019, 11, 364
117. Garrido, D.; Ruiz-Moyano, S.; Lemay, D.G.; Sela, D.A.; German, J.B.; Mills, D.A. Comparative transcriptomics
reveals key differences in the response to milk oligosaccharides of infant gut-associated bifidobacteria.
Sci. Rep. 2015, 5, 13517. [CrossRef] [PubMed]
118. Ruiz-Moyano, S.; Totten, S.M.; Garrido, D.; Smilowitz, J.T.; German, J.B.; Lebrilla, C.B.; Mills, D.A. Variation
in consumption of human milk oligosaccharides by infant-gut associated strains of bifidobacterium breve.
Appl. Environ. Microbiol. 2013, 79, 6040–6049. [CrossRef] [PubMed]
119. González, R.; Klaassens, E.S.; Malinen, E.; De Vos, W.M.; Vaughan, E.E. Differential transcriptional response
of bifidobacterium longum to human milk, formula milk, and galactooligosaccharide. Appl. Environ.
Microbiol. 2008, 74, 4686–4694. [CrossRef] [PubMed]
120. Hugenholtz, F.; Ritari, J.; Nylund, L.; Davids, M.; Satokari, R.; De Vos, W.M. Feasibility of metatranscriptome
analysis from infant gut microbiota: Adaptation to solid foods results in increased activity of firmicutes at
six months. Int. J. Microbiol. 2017, 2017. [CrossRef] [PubMed]
121. Fallani, M.; Amarri, S.; Uusijarvi, A.; Adam, R.; Khanna, S.; Aguilera, M.; Gil, A.; Vieites, J.M.; Norin, E.;
Young, D.; et al. Determinants of the human infant intestinal microbiota after the introduction of first
complementary foods in infant samples from five european centres. Microbiology 2011, 157, 1385–1392.
[CrossRef] [PubMed]
122. Thompson, A.L.; Monteagudo-Mera, A.; Cadenas, M.B.; Lampl, M.L.; Azcarate-Peril, M.A. Milk- and
solid-feeding practices and daycare attendance are associated with differences in bacterial diversity,
predominant communities, and metabolic and immune function of the infant gut microbiome. Front. Cell.
Infect. Microbiol. 2015, 5, 3. [CrossRef] [PubMed]
123. Scheiwiller, J.; Arrigoni, E.; Brouns, F.; Amado, R. Human faecal microbiota develops the ability to degrade
type 3 resistant starch during weaning. J. Pediatr. Gastroenterol. Nutr. 2006, 43, 584–591. [CrossRef]
124. Birt, D.F.; Boylston, T.; Hendrich, S.; Jane, J.L.; Hollis, J.; Li, L.; McClelland, J.; Moore, S.; Phillips, G.J.;
Rowling, M.; et al. Resistant starch: Promise for improving human health. Adv. Nutr. 2013, 4, 587–601.
[CrossRef] [PubMed]
125. Flint, H.J.; Scott, K.P.; Duncan, S.H.; Louis, P.; Forano, E. Microbial degradation of complex carbohydrates in
the gut. Gut Microbes 2012, 3, 289–306. [CrossRef] [PubMed]
126. Kovatcheva-Datchary, P.; Egert, M.; Maathuis, A.; Rajilić-Stojanović, M.; De Graaf, A.A.; Smidt, H.; De
Vos, W.M.; Venema, K. Linking phylogenetic identities of bacteria to starch fermentation in an in vitro model
of the large intestine by rna-based stable isotope probing. Environ. Microbiol. 2009, 11, 914–926. [CrossRef]
[PubMed]
127. Shipman, J.A.; Berleman, J.E.; Salyers, A.A. Characterization of four outer membrane proteins involved in
binding starch to the cell surface of bacteroides thetaiotaomicron. J. Bacteriol. 2000, 182, 5365–5372. [CrossRef]
128. Crittenden, R.; Laitila, A.; Forssell, P.; Matto, J.; Saarela, M.; Mattila-Sandholm, T.; Myllarinen, P. Adhesion
of bifidobacteria to granular starch and its implications in probiotic technologies. Appl. Environ. Microbiol.
2001, 67, 3469–3475. [CrossRef]
129. Shipman, J.A.; Cho, K.H.; Siegel, H.A.; Salyers, A.A. Physiological characterization of susg, an outer
membrane protein essential for starch utilization by bacteroides thetaiotaomicron. J. Bacteriol. 1999, 181,
7206–7211. [PubMed]
130. Koropatkin, N.M.; Smith, T.J. Susg: A unique cell-membrane-associated alpha-amylase from a prominent
human gut symbiont targets complex starch molecules. Structure 2010, 18, 200–215. [CrossRef]
131. Stewart, M.L.; Timm, D.A.; Slavin, J.L. Fructooligosaccharides exhibit more rapid fermentation than
long-chain inulin in an in vitro fermentation system. Nutr. Res. 2008, 28, 329–334. [CrossRef] [PubMed]
132. Warren, F.J.; Fukuma, N.M.; Mikkelsen, D.; Flanagan, B.M.; Williams, B.A.; Lisle, A.T.; Cuív, P.Ó.;
Morrison, M.; Gidley, M.J. Food starch structure impacts gut microbiome composition. mSphere 2018,
3, e00086-18. [CrossRef]
133. Wong, J.M.; de Souza, R.; Kendall, C.W.; Emam, A.; Jenkins, D.J. Colonic health: Fermentation and short
chain fatty acids. J. Clin. Gastroenterol. 2006, 40, 235–243. [CrossRef]
134. Morrison, D.J.; Preston, T. Formation of short chain fatty acids by the gut microbiota and their impact on
human metabolism. Gut Microbes 2016, 7, 189–200. [CrossRef] [PubMed]
135. Herrmann, E.; Young, W.; Rosendale, D.; Conrad, R.; Riedel, C.U.; Egert, M. Determination of resistant starch
assimilating bacteria in fecal samples of mice by in vitro rna-based stable isotope probing. Front. Microbiol.
2017, 8, 1331. [CrossRef]
75
Nutrients 2019, 11, 364
136. Boets, E.; Gomand, S.V.; Deroover, L.; Preston, T.; Vermeulen, K.; Preter, V.; Hamer, H.M.; den Mooter, G.;
Vuyst, L.; Courtin, C.M. Systemic availability and metabolism of colonic-derived short-chain fatty acids in
healthy subjects: A stable isotope study. J. Physiol. 2017, 595, 541–555. [CrossRef]
137. Koga, Y.; Tokunaga, S.; Nagano, J.; Sato, F.; Konishi, K.; Tochio, T.; Murakami, Y.; Masumoto, N.; Tezuka, J.-I.;
Sudo, N.; et al. Age-associated effect of kestose on faecalibacterium prausnitzii and symptoms in the atopic
dermatitis infants. Pediatr. Res. 2016, 80, 844–851. [CrossRef] [PubMed]
138. Bridgman, S.L.; Azad, M.B.; Field, C.J.; Haqq, A.M.; Becker, A.B.; Mandhane, P.J.; Subbarao, P.; Turvey, S.E.;
Sears, M.R.; Scott, J.A.; et al. Fecal short-chain fatty acid variations by breastfeeding status in infants at 4
months: Differences in relative versus absolute concentrations. Front. Nutr. 2017, 4, 11. [CrossRef] [PubMed]
139. Macfarlane, S.; Macfarlane, G.T. Regulation of short-chain fatty acid production. Proc. Nutr. Soc. 2003, 62,
67–72. [CrossRef]
140. Vinolo, M.A.; Rodrigues, H.G.; Nachbar, R.T.; Curi, R. Regulation of inflammation by short chain fatty acids.
Nutrients 2011, 3, 858–876. [CrossRef]
141. Kalina, U.; Koyama, N.; Hosoda, T.; Nuernberger, H.; Sato, K.; Hoelzer, D.; Herweck, F.; Manigold, T.;
Singer, M.V.; Rossol, S. Enhanced production of il-18 in butyrate-treated intestinal epithelium by stimulation
of the proximal promoter region. Eur. J. Immunol. 2002, 32, 2635–2643. [CrossRef]
142. Louis, P.; Flint, H.J. Formation of propionate and butyrate by the human colonic microbiota. Environ. Microbiol.
2017, 19, 29–41. [CrossRef] [PubMed]
143. Tedelind, S.; Westberg, F.; Kjerrulf, M.; Vidal, A. Anti-inflammatory properties of the short-chain fatty acids
acetate and propionate: A study with relevance to inflammatory bowel disease. World J. Gastroenterol. 2007,
13, 2826. [CrossRef] [PubMed]
144. Fukuda, S.; Toh, H.; Taylor, T.D.; Ohno, H.; Hattori, M. Acetate-producing bifidobacteria protect the host
from enteropathogenic infection via carbohydrate transporters. Gut Microbes 2012, 3, 449–454. [CrossRef]
145. Hosseini, E.; Grootaert, C.; Verstraete, W.; Van de Wiele, T. Propionate as a health-promoting microbial
metabolite in the human gut. Nutr. Rev. 2011, 69, 245–258. [CrossRef]
146. Mariadason, J.M.; Barkla, D.H.; Gibson, P.R. Effect of short-chain fatty acids on paracellular permeability in
caco-2 intestinal epithelium model. Am. J. Physiol. 1997, 272, G705–G712. [CrossRef]
147. Gantois, I.; Ducatelle, R.; Pasmans, F.; Haesebrouck, F.; Hautefort, I.; Thompson, A.; Hinton, J.C.; Van
Immerseel, F. Butyrate specifically down-regulates salmonella pathogenicity island 1 gene expression.
Appl. Environ. Microbiol. 2006, 72, 946–949. [CrossRef]
148. Louis, P.; Flint, H.J. Diversity, metabolism and microbial ecology of butyrate-producing bacteria from the
human large intestine. FEMS Microbiol. Lett. 2009, 294, 1–8. [CrossRef]
149. Peng, L.; Li, Z.-R.; Green, R.S.; Holzman, I.R.; Lin, J. Butyrate enhances the intestinal barrier by facilitating
tight junction assembly via activation of amp-activated protein kinase in caco-2 cell monolayers. J. Nutr.
2009, 139, 1619–1625. [CrossRef]
150. Yatsunenko, T.; Rey, F.E.; Manary, M.J.; Trehan, I.; Dominguez-Bello, M.G.; Contreras, M.; Magris, M.;
Hidalgo, G.; Baldassano, R.N.; Anokhin, A.P. Human gut microbiome viewed across age and geography.
Nature 2012, 486, 222. [CrossRef] [PubMed]
151. Magnúsdóttir, S.; Ravcheev, D.; de Crécy-Lagard, V.; Thiele, I. Systematic genome assessment of b-vitamin
biosynthesis suggests co-operation among gut microbes. Front. Genet. 2015, 6, 148. [CrossRef]
152. Biesalski, H.K. Nutrition meets the microbiome: Micronutrients and the microbiota. Ann. N. Y. Acad. Sci.
2016, 1372, 53–64. [CrossRef] [PubMed]
153. Ueland, P.M.; McCann, A.; Midttun, Ø.; Ulvik, A. Inflammation, vitamin b6 and related pathways. Mol.
Aspects Med. 2017, 53, 10–27. [CrossRef] [PubMed]
154. Kunisawa, J.; Hashimoto, E.; Ishikawa, I.; Kiyono, H. A pivotal role of vitamin b9 in the maintenance of
regulatory t cells in vitro and in vivo. PloS ONE 2012, 7, e32094. [CrossRef] [PubMed]
155. Tamura, J.; Kubota, K.; Murakami, H.; Sawamura, M.; Matsushima, T.; Tamura, T.; Saitoh, T.; Kurabayshi, H.;
Naruse, T. Immunomodulation by vitamin B12: Augmentation of CD8+ T lymphocytes and natural killer
(nk) cell activity in vitamin B12-deficient patients by methyl-B12 treatment. Clin. Exp. Immunol. 1999, 116,
28–32. [CrossRef]
156. Kjer-Nielsen, L.; Patel, O.; Corbett, A.J.; Le Nours, J.; Meehan, B.; Liu, L.; Bhati, M.; Chen, Z.; Kostenko, L.;
Reantragoon, R. Mr1 presents microbial vitamin b metabolites to mait cells. Nature 2012, 491, 717. [CrossRef]
[PubMed]
76
Nutrients 2019, 11, 364
157. de Verteuil, D.; Granados, D.P.; Thibault, P.; Perreault, C. Origin and plasticity of mhc i-associated self
peptides. Autoimmun. Rev. 2012, 11, 627–635. [CrossRef] [PubMed]
158. Gibson, R.S.; Bailey, K.B.; Gibbs, M.; Ferguson, E.L. A review of phytate, iron, zinc, and calcium
concentrations in plant-based complementary foods used in low-income countries and implications for
bioavailability. Food Nutr. Bull. 2010, 31, S134–S146. [CrossRef] [PubMed]
© 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
77
nutrients
Review
Vitamin D: Nutrient, Hormone,
and Immunomodulator
Francesca Sassi, Cristina Tamone and Patrizia D’Amelio *
Department of Medical Science, Gerontology and Bone Metabolic Diseases, University of Turin,
10126 Turin, Italy; francesca.sassi@unito.it (F.S.); cristinatamone78@gmail.com (C.T.)
* Correspondence: patrizia.damelio@unito.it; Tel.: +39-011-6335533
Abstract: The classical functions of vitamin D are to regulate calcium-phosphorus homeostasis and
control bone metabolism. However, vitamin D deficiency has been reported in several chronic
conditions associated with increased inflammation and deregulation of the immune system, such as
diabetes, asthma, and rheumatoid arthritis. These observations, together with experimental studies,
suggest a critical role for vitamin D in the modulation of immune function. This leads to the
hypothesis of a disease-specific alteration of vitamin D metabolism and reinforces the role of
vitamin D in maintaining a healthy immune system. Two key observations validate this important
non-classical action of vitamin D: first, vitamin D receptor (VDR) is expressed by the majority
of immune cells, including B and T lymphocytes, monocytes, macrophages, and dendritic cells;
second, there is an active vitamin D metabolism by immune cells that is able to locally convert
25(OH)D3 into 1,25(OH)2 D3 , its active form. Vitamin D and VDR signaling together have a
suppressive role on autoimmunity and an anti-inflammatory effect, promoting dendritic cell and
regulatory T-cell differentiation and reducing T helper Th 17 cell response and inflammatory cytokines
secretion. This review summarizes experimental data and clinical observations on the potential
immunomodulating properties of vitamin D.
1. Introduction
The role of vitamin D in the regulation of calcium-phosphate homeostasis and in the control of
bone turnover is well known. Vitamin D status significantly affects skeletal health during growth and
in adult age, its deficiency during growth leads to rickets [1], whereas during adult age it is responsible
of osteomalacia and various degree of osteoporo-malacia [2]. Low vitamin D status increases bone
turnover, decreases bone density, and is associated with increased fracture risk. In addition to the
well-known effect on skeletal health in the last two decades evidence has been accumulated on the
pleiotropic effect of vitamin D other than on bone health thanks to the findings that vitamin D receptor
(VDR) and the vitamin D activating enzyme 1-α-hydroxylase (CYP27B1) are expressed in several cells
outside the bone and kidney, such as in the intestine, platelets, pancreas, and prostate [3]. Several cells
involved in the immune function express VDR and CYP27B1, this observation suggests that the active
form of vitamin D, 1,25(OH)2 D3 , is able to control the immune function at different levels. Previous
reviews on the role of vitamin D in the regulation of immune system have been published in recent
years [4,5]. Here we summarize the recent evidence sexploiting authors’ expertise in both experimental
and clinical fields.
2. Vitamin D Metabolism
Vitamin D enters the body trough dietary intake (about 20% of vitamin D3 is assumed with diet)
or is synthetized by the skin (80%) from 7-dihydrocholesterol following UVB exposure. Vitamin D3
becomes biologically active after hydroxylation in the liver by the enzymes cytochrome P450 2R1
(CYP2R1) and cytochrome P450 27 (CYP27A1) becoming 25(OH)D3 . The fully-active metabolite
1,25(OH)2 D3 is hydroxylated in the kidney by the enzyme CYP27B1, parathormone (PTH), and the
fibroblast growth factor 23 (FGF-23) control CYP27B1 synthesis and activity [6]. Synthesis of
1,25(OH)2 D3 is strictly regulated in a renal negative feedback loop: high levels of 1,25(OH)2 D3
and FGF-23 inhibit CYP27B1 and induce the cytochrome P45024A1(CYP24A1), which transforms
1,25(OH)2 D3 into the inactive form 24(OH)D3 [7].
In addition to the kidney, CYP27B1 is expressed by other cell types, including immune cells.
These cells produce 1,25(OH)2 D3 that has autocrine and/or paracrine effects, the high level produced
locally is thought to be responsible for immunomodulation. The regulation of CYP27B1 synthesis in
immune cells is different than the signals regulating kidney production of 1,25(OH)2 D3 . Inflammatory
signals, such as lipopolysaccharide (LPS) and cytokines, induce monocyte and macrophage production
of CYP27B1 [8–10]. These differences in the regulation of 1,25(OH)2 D3 production point to an
autocrine/paracrine effect as immunomodulatory.
3. Vitamin D Status
Vitamin D status is defined by the blood measurement of its hydroxylated form 25(OH)D3 ,
however, there is no common agreement on the threshold levels to identify desirable vitamin D level.
Guidelines from different scientific societies and different countries established 50 nM/L or 75 nM/L
to consider vitamin D sufficiency [11–13], however, it is generally accepted that 25(OH)D3 levels lower
than 50 nM/L are associated with bone metabolism alteration, increased risk of falls, and myopathy in
adults [14–18]. Experts in the field generally agree to maintain 25(OH)D3 between 20 and 125 nM/L in
order to obtain the certain skeletal effects without toxic effects. Recent literature raises the suspicion
that administration of a bolus of vitamin D3 higher than 50,000 UI may result in an increased risk of
falls and fractures [19,20]; moreover, the mortality related to 25(OH)D3 is a “U shaped curve” and
25(OH)D3 levels higher than 150 nM/L are associated with increased mortality [21].
79
Nutrients 2018, 10, 1656
1,25(OH)2 D3 up-regulates CAMP not only by monocytes/macrophages, but also in other cells
participating in the innate immune system as first-barrier defenses, such as keratinocytes, epithelial,
intestinal, lung and corneal cells, and placenta trophoblasts (see for a comprehensive review Wei and
Christakos, 2015) [4].
Data in humans on infections other than mycobacterial have been generated on urinary and
respiratory infections and on sepsis. A predisposition to urinary tract infection in children with low
vitamin D levels due to the reduced production of CAMP and defensing β2 has been suggested
by association studies [31,32]. Additionally, in chronic obstructive pulmonary disease (COPD)
patients’ levels of CAMP and other antimicrobial peptides were associated with increased risk of acute
exacerbations [33]. Consistent with this datum treatment with 1,25(OH)2 D3 was effective in reducing
respiratory infections in asthma patients thanks to increased CAMP expression and inflammatory
cytokine modulation [34]. Data on the role of vitamin D status and vitamin D supplementation in
sepsis are also available both in pediatric and in adult patients: in pediatric patients a clear role
for 25(OH)D3 and CAMP was not demonstrated [35], whereas in adults lower levels of 25(OH)D3
were found in sepsis [36] and a high-dose of vitamin D3 increases circulating CAMP and reduces
inflammatory cytokines as IL-6 and IL-1β [37].
More recently data on a possible role of vitamin D in increasing resistance to HIV infection
have been published, in particular HIV-exposed seronegative individuals produced more CAMP in
oral-mucosa and peripheral-blood, and have higher CYP24A1 mRNA in vaginal-mucosa; CYP24A1 is
considered an indicator of high levels of 1,25(OH)2 D3 [38]. Low serum vitamin D has been associated
with HIV/AIDS progression and mortality [39].
1,25(OH)2 D3 is able to increase the production of other antimicrobial peptides, such as defensing
β2-4, this ability has been demonstrated both in vitro by monocytes stimulation [40,41] and in vivo in
pediatric patients’ blood [32].
Vitamin D is able to modulate innate immune system, also increasing the phagocytic ability on
immune cells [42,43] and by reinforcing the physical barrier function of epithelial cells. In particular
1,25(OH)2 D3 can enhance corneal [44] and intestinal [45] epithelial barrier function (Figure 1).
Taken together these data point to a role of vitamin D in defending the organism against pathogens
suggesting that vitamin D sufficiency has to be granted in patients affected by acute or chronic
infection. The ability of immune cells to hydroxylate 25(OH)D3 into its active form 1,25(OH)2 D3
suggests administrating vitamin D3 rather than hydroxylated metabolites to patients affected by
infections in order to allow the autocrine/paracrine function of 1,25(OH)2 D3 without overcoming local
hydroxylation and the feedback system.
80
Nutrients 2018, 10, 1656
epithelial cell apoptosis, boosting tight junction function [51,53]. On the other hand VDR selective
deletion in bowel favors a more severe form of colitis characterized by greater Th1 and Th17 mucosal
infiltration and inflammatory cytokines production [54]. In humans, observational studies suggest that
low levels of 25(OH)D3 are associated with increased risk of inflammatory bowel disease (IBD) [55–57]
and that high levels of 25(OH)D3 in these patients protect against Clostridium difficile infection [58].
The experimental data on the role of VDR in developing IBD have been confirmed by the finding of a
significant reduction of VDR expression (about 50%) in the colon epithelium in patients affected by IBD
with respect to healthy controls [51,53]. The reduction in VDR expression by IBD patients may explain
the different effect on GM composition of high oral dosages of vitamin D3 demonstrated in a small
cohort of patients affected by Crohn’s disease with respect to healthy controls [59], however, human
data on the effect of vitamin D supplementation on GM in IBD are still controversial, as other studies
did not confirm these results [60,61]. In the study by Luthold and coll. [61] dietary intake of vitamin D
and 25(OH)D3 were inversely correlated with Coprococcus and Bifidobacterium, however, thanks to their
ability to produce butyrate these bacteria are commonly considered as anti-inflammatory. A possible
explanation of these contradictory results may be the different effect of vitamin D on GM according
to the different gastro-intestinal tracts considered [62]. Recently, a double-blind placebo-controlled
study on patients affected by cystic fibrosis demonstrated that vitamin D insufficiency is associated
with different microbiota not only in the gut, but also in the airways, and that the administration of
50,000 IU of oral vitamin D3 weekly significantly affects microbiota composition [63]. Nevertheless,
the evaluation of clinical outcomes of microbiota change is still open.
Several data point to an effect of vitamin D on microbiota. Conversely, some recent reports suggest
that microbiota, per se, influences vitamin D metabolism mainly through FGF-23; germ-free (GF) mice
have low vitamin D and high FGF-23, whereas their colonization with bacteria results in increased
levels of tumor necrosis factor-α (TNF-α) and a decrease in FGF-23 with normalization of vitamin D
hydroxylated metabolites. Inhibition of FGF-23 in GF mice restores vitamin D metabolism without
bacterial colonization of the gut [64] (Figure 1).
The role of GM as an active player in the regulation of bone metabolism in humans is being
investigated more and more [46], and the role played by vitamin D is still under debate. Further studies
to clarify their interplay are needed.
81
Nutrients 2018, 10, 1656
Figure 1. Effects of vitamin D on the innate immune system and gut microbiota. Abbreviations: EC,
enteral cells; GM, gut microbiota.
82
Nutrients 2018, 10, 1656
(RUNx1) binding to the IL-17 promoter and inducing Forkhead box P3 (FOXP3) [81], and by inhibiting
RAR-related Orphan Receptor Gamma2 (RORγt) which is the transcription factor of IL-17 [84].
More recently our lab showed no effect of the administration of a high bolus of vitamin D3
(300,000 UI) in the modulation of Th subset in patients affected by early rheumatoid arthritis [73],
as well as a study on hemodialysis patients [72].
It has also been suggested that the administration of oral vitamin D3 increases Tregs function
in patients with type 1 diabetes mellitus [85], however, in other diseases, such as early rheumatoid
arthritis, this effect was not confirmed [73].
The overall effect of vitamin D on Th cells differentiation may be mediated by its effect on
dendritic cells, these cells are antigen-presenting cells (APCs), responsible for T cell differentiation
into an effector cell with pro- or anti-inflammatory properties, thus, modulation of APCs is crucial in
initiating and maintaining adaptive immune response and self-tolerance [86]. In vitro differentiation of
dendritic cells in the presence of 1,25(OH)2 D3 induces a “tolerogenic state” characterized by low levels
of inflammatory cytokines, such as IL-12 and TNF-α, with increased levels of the anti-inflammatory
IL-10, these cells induce the differentiation of Treg cells and induce apoptosis in the autoreactive T
cells [87–90] (Figure 2).
Taken together these data are not sufficient to prove a real role for vitamin D in the modulation
of adaptive immune system in humans, thus, the therapeutic use of vitamin D and its metabolites in
patients aiming to ameliorate the adaptive immune system is not sustained by sufficient data.
Figure 2. Effect of vitamin D on the adaptive immune system. Abbreviations: APC, antigen presenting
cell; IFN, interferon; IL, interleukin; Th1, T helper 1 cell; Th2, T helper 2 cell; Th17, T helper 17 cell;
TNF, tumor necrosis factor; Treg, T regulatory cell.
83
Nutrients 2018, 10, 1656
8. Conclusions
In summary, several studies point to an important role of vitamin D as an immunomodulator,
and strong data demonstrate a role for 1,25(OH)2 D3 in increasing the ability of the innate immune
system to fight against pathogens, whereas data on the effect of 1,25(OH)2 D3 in the modulation of
acquired immune system are more controversial. There is no general consensus on the desired level of
84
Nutrients 2018, 10, 1656
25(OH)D3 to achieve immunomodulatory effects, thus, there is no current indication for vitamin D3
supplementation in patients with infections and/or autoimmune diseases. Further studies are needed
to clarify the role of vitamin D as immunomodulator in humans.
Author Contributions: All three authors participated in the bibliographic search, discussion and writing of the
manuscript. The manuscript was finalized by P.D.
Funding: This research received no external funding
Conflicts of Interest: The authors declare that they have no conflict of interest.
References
1. Antonucci, R.; Locci, C.; Clemente, M.G.; Chicconi, E.; Antonucci, L. Vitamin D deficiency in childhood: Old
lessons and current challenges. J. Pediatr. Endocrinol. Metab. 2018, 31, 247–260. [CrossRef] [PubMed]
2. Uday, S.; Högler, W. Prevention of rickets and osteomalacia in the UK: Political action overdue.
Arch. Dis. Child. 2018, 103, 901–906. [CrossRef] [PubMed]
3. D’Amelio, P.; Cristofaro, M.A.; De Vivo, E.; Ravazzoli, M.; Grosso, E.; Di Bella, S.; Aime, M.; Cotto, N.;
Silvagno, F.; Isaia, G.; et al. Platelet vitamin D receptor is reduced in osteoporotic patients. Panminerva Med.
2012, 54, 225–231. [PubMed]
4. Wei, R.; Christakos, S. Mechanisms Underlying the regulation of innate and adaptive immunity by vitamin
D. Nutrients 2015, 7, 8251–8260. [CrossRef] [PubMed]
5. Altieri, B.; Muscogiuri, G.; Barrea, L.; Mathieu, C.; Vallone, C.V.; Mascitelli, L.; Bizzaro, G.; Altieri, V.M.;
Tirabassi, G.; Balercia, G.; et al. Does vitamin D play a role in autoimmune endocrine? A proof of concept.
Rev. Endocr. Metab. Disord. 2013, 18, 335–346. [CrossRef] [PubMed]
6. Borel, P.; Caillaud, D.; Cano, N.J. Vitamin D bioavailability: State of the art. Crit. Rev. Food Sci. Nutr. 2015, 55,
1193–1205. [CrossRef] [PubMed]
7. Anderson, P.H. Vitamin D Activity and Metabolism in Bone. Curr. Osteoporos. Rep. 2017, 15, 443–449.
[CrossRef] [PubMed]
8. Liu, P.T.; Stenger, S.; Li, H.; Wenzel, L.; Tan, B.H.; Krutzik, S.R.; Ochoa, M.T.; Schauber, J.; Wu, K.;
Meinken, C.; et al. Toll-like receptor triggering of a vitamin D-mediated human antimicrobial response.
Science 2006, 311, 1770–1773. [CrossRef] [PubMed]
9. Stoffels, K.; Overbergh, L.; Giulietti, A.; Verlinden, L.; Bouillon, R.; Mathieu, C. Immune regulation of
25-hydroxyvitamin-D3 -1alpha-hydroxylase in human monocytes. J. Bone Miner. Res. 2006, 21, 37–47.
[CrossRef] [PubMed]
10. Krutzik, S.R.; Hewison, M.; Liu, P.T.; Robles, J.A.; Stenger, S.; Adams, J.S.; Modlin, R.L. IL-15 links
TLR2/1-induced macrophage differentiation to the vitamin D-dependent antimicrobial pathway. J. Immunol.
2008, 181, 7115–7120. [CrossRef] [PubMed]
11. Ross, A.C.; Manson, J.E.; Abrams, S.A.; Aloia, J.F.; Brannon, P.M.; Clinton, S.K.; Durazo-Arvizu, R.A.;
Gallagher, J.C.; Gallo, R.L.; Jones, G.; et al. The 2011 report on dietary reference intakes for calcium and
vitamin D from the Institute of Medicine: What clinicians need to know. J. Clin. Endocrinol. Metab. 2011, 96,
53–58. [CrossRef] [PubMed]
12. Holick, M.F.; Binkley, N.C.; Bischoff-Ferrari, H.A.; Gordon, C.M.; Hanley, D.A.; Heaney, R.P.; Murad, M.H.;
Weaver, C.M. Endocrine Society Evaluation, treatment, and prevention of vitamin D deficiency: An Endocrine
Society clinical practice guideline. J. Clin. Endocrinol. Metab. 2011, 96, 1911–1930. [CrossRef] [PubMed]
13. Cesareo, R.; Attanasio, R.; Caputo, M.; Castello, R.; Chiodini, I.; Falchetti, A.; Guglielmi, R.; Papini, E.;
Santonati, A.; Scillitani, A.; et al. Italian Association of Clinical Endocrinologists (AME) and Italian chapter
of the American Association of Clinical Endocrinologists (AACE) position statement: Clinical management
of vitamin D deficiency in adults. Nutrients 2018, 10, 546. [CrossRef] [PubMed]
14. Valcour, A.; Blocki, F.; Hawkins, D.M.; Rao, S.D. Effects of age and serum 25-OH-vitamin D on serum
parathyroid hormone levels. J. Clin. Endocrinol. Metab. 2012, 97, 3989–3995. [CrossRef] [PubMed]
15. LeBlanc, E.; Chou, R.; Zakher, B.; Daeges, M.; Pappas, M. Screening for Vitamin D Deficiency: Systematic
Review for the U.S. Preventive Services Task Force Recommendation; Agency for Healthcare Research and Quality:
Rockville, MD, USA, 2014.
85
Nutrients 2018, 10, 1656
16. LeBlanc, E.S.; Zakher, B.; Daeges, M.; Pappas, M.; Chou, R. Screening for vitamin D deficiency: A systematic
review for the U.S. Preventive Services Task Force. Ann. Intern. Med. 2015, 162, 109–122. [CrossRef]
[PubMed]
17. Gillespie, L.D.; Robertson, M.C.; Gillespie, W.J.; Sherrington, C.; Gates, S.; Clemson, L.M.; Lamb, S.E.
Interventions for preventing falls in older people living in the community. Cochrane Database Syst. Rev. 2012.
[CrossRef] [PubMed]
18. Bhattoa, H.P.; Konstantynowicz, J.; Laszcz, N.; Wojcik, M.; Pludowski, P. Vitamin D: Musculoskeletal health.
Rev. Endocr. Metab. Disord. 2017, 18, 363–371. [CrossRef] [PubMed]
19. Sanders, K.M.; Stuart, A.L.; Williamson, E.J.; Simpson, J.A.; Kotowicz, M.A.; Young, D.; Nicholson, G.C.
Annual high-dose oral vitamin D and falls and fractures in older women: A randomized controlled trial.
JAMA 2010, 303, 1815–1822. [CrossRef] [PubMed]
20. Bischoff-Ferrari, H.A.; Dawson-Hughes, B.; Orav, E.J.; Staehelin, H.B.; Meyer, O.W.; Theiler, R.; Dick, W.;
Willett, W.C.; Egli, A. Monthly high-dose vitamin D Treatment for the prevention of functional decline: A
randomized clinical trial. JAMA Intern. Med. 2016, 176, 175–183. [CrossRef] [PubMed]
21. Amrein, K.; Quraishi, S.A.; Litonjua, A.A.; Gibbons, F.K.; Pieber, T.R.; Camargo, C.A.; Giovannucci, E.;
Christopher, K.B. Evidence for a U-shaped relationship between prehospital vitamin D status and mortality:
A cohort study. J. Clin. Endocrinol. Metab. 2014, 99, 1461–1469. [CrossRef] [PubMed]
22. Airey, F.S. Vitamin D as a remedy for lupus vulgaris. Med. World 1946, 64, 807–810. [PubMed]
23. Herrera, G. Vitamin D in massive doses as an adjuvant to the sulfones in the treatment of tuberculoid leprosy.
Int. J. Lepr. 1949, 17, 35–42. [PubMed]
24. Wang, T.-T.; Nestel, F.P.; Bourdeau, V.; Nagai, Y.; Wang, Q.; Liao, J.; Tavera-Mendoza, L.; Lin, R.;
Hanrahan, J.W.; Mader, S.; et al. Cutting edge: 1,25-dihydroxyvitamin D3 is a direct inducer of antimicrobial
peptide gene expression. J. Immunol. 2004, 173, 2909–2912. [CrossRef] [PubMed]
25. Gombart, A.F.; Borregaard, N.; Koeffler, H.P. Human cathelicidin antimicrobial peptide (CAMP) gene is a
direct target of the vitamin D receptor and is strongly up-regulated in myeloid cells by 1,25-dihydroxyvitamin
D3 . FASEB J. 2005, 19, 1067–1077. [CrossRef] [PubMed]
26. Dai, X.; Sayama, K.; Tohyama, M.; Shirakata, Y.; Hanakawa, Y.; Tokumaru, S.; Yang, L.; Hirakawa, S.;
Hashimoto, K. PPARγ mediates innate immunity by regulating the 1α,25-dihydroxyvitamin D3 induced
hBD-3 and cathelicidin in human keratinocytes. J. Dermatol. Sci. 2010, 60, 179–186. [CrossRef] [PubMed]
27. Liu, P.T.; Stenger, S.; Tang, D.H.; Modlin, R.L. Cutting edge: Vitamin D-mediated human antimicrobial
activity against Mycobacterium tuberculosis is dependent on the induction of cathelicidin. J. Immunol. 2007,
179, 2060–2063. [CrossRef] [PubMed]
28. White, J.H. Vitamin D as an inducer of cathelicidin antimicrobial peptide expression: Past, present and
future. J. Steroid Biochem. Mol. Biol. 2010, 121, 234–238. [CrossRef] [PubMed]
29. Bursuker, I.; Goldman, R. On the origin of macrophage heterogeneity: A hypothesis. J. Reticuloendothel. Soc.
1983, 33, 207–220. [PubMed]
30. Kim, E.W.; Teles, R.M.B.; Haile, S.; Liu, P.T.; Modlin, R.L. Vitamin D status contributes to the antimicrobial
activity of macrophages against Mycobacterium leprae. PLoS Negl. Trop. Dis. 2018, 12. [CrossRef] [PubMed]
31. ÖvünçHacıhamdioğlu, D.; Altun, D.; Hacıhamdioğlu, B.; Çekmez, F.; Aydemir, G.; Kul, M.; Müftüoğlu, T.;
Süleymanoğlu, S.; Karademir, F. The association between serum 25-hydroxy vitamin D level and urine
cathelicidin in children with a urinary tract infection. J. Clin. Res. Pediatr. Endocrinol. 2016, 8, 325–329.
[CrossRef] [PubMed]
32. Georgieva, V.; Kamolvit, W.; Herthelius, M.; Lüthje, P.; Brauner, A.; Chromek, M. Association
between vitamin D, antimicrobial peptides and urinary tract infection in infants and young children.
Acta Paediatr. 2018. [CrossRef] [PubMed]
33. Persson, L.J.P.; Aanerud, M.; Hardie, J.A.; Miodini Nilsen, R.; Bakke, P.S.; Eagan, T.M.; Hiemstra, P.S.
Antimicrobial peptide levels are linked to airway inflammation, bacterial colonisation and exacerbations in
chronic obstructive pulmonary disease. Eur. Respir. J. 2017, 49. [CrossRef] [PubMed]
34. Ramos-Martínez, E.; López-Vancell, M.R.; Fernández de Córdova-Aguirre, J.C.; Rojas-Serrano, J.;
Chavarría, A.; Velasco-Medina, A.; Velázquez-Sámano, G. Reduction of respiratory infections in asthma
patients supplemented with vitamin D is related to increased serum IL-10 and IFNγ levels and cathelicidin
expression. Cytokine 2018, 108, 239–246. [CrossRef] [PubMed]
86
Nutrients 2018, 10, 1656
35. Mathias, E.; Tangpricha, V.; Sarnaik, A.; Farooqi, A.; Sethuraman, U. Association of vitamin D with
cathelicidin and vitamin D binding protein in pediatric sepsis. J. Clin. Transl. Endocrinol. 2017, 10, 36–38.
[CrossRef] [PubMed]
36. Greulich, T.; Regner, W.; Branscheidt, M.; Herr, C.; Koczulla, A.R.; Vogelmeier, C.F.; Bals, R. Altered
blood levels of vitamin D, cathelicidin and parathyroid hormone in patients with sepsis-a pilot study.
Anaesth. Intensive Care 2017, 45, 36–45. [PubMed]
37. Quraishi, S.A.; De Pascale, G.; Needleman, J.S.; Nakazawa, H.; Kaneki, M.; Bajwa, E.K.; Camargo, C.A.;
Bhan, I. Effect of cholecalciferol supplementation on vitamin D status and cathelicidin levels in sepsis: A
randomized, placebo-controlled trial. Crit. Care Med. 2015, 43, 1928–1937. [CrossRef] [PubMed]
38. Aguilar-Jimenez, W.; Zapata, W.; Rugeles, M.T. Antiviral molecules correlate with vitamin D pathway genes
and are associated with natural resistance to HIV-1 infection. Microbes. Infect. 2016, 18, 510–516. [CrossRef]
[PubMed]
39. Coussens, A.K.; Naude, C.E.; Goliath, R.; Chaplin, G.; Wilkinson, R.J.; Jablonski, N.G. High-dose vitamin D3
reduces deficiency caused by low UVB exposure and limits HIV-1 replication in urban Southern Africans.
PNAS 2015, 112, 8052–8057. [CrossRef] [PubMed]
40. Liu, P.T.; Schenk, M.; Walker, V.P.; Dempsey, P.W.; Kanchanapoomi, M.; Wheelwright, M.; Vazirnia, A.;
Zhang, X.; Steinmeyer, A.; Zügel, U.; et al. Convergence of IL-1beta and VDR activation pathways in human
TLR2/1-induced antimicrobial responses. PLoS ONE 2009, 4. [CrossRef]
41. Castañeda-Delgado, J.E.; Araujo, Z.; Gonzalez-Curiel, I.; Serrano, C.J.; Rivas Santiago, C.; Enciso-Moreno, J.A.;
Rivas-Santiago, B. Vitamin D and l-isoleucine promote antimicrobial peptide hBD-2 production in peripheral
blood mononuclear cells from elderly individuals. Int. J. Vitam. Nutr. Res. 2016, 86, 56–61. [CrossRef]
[PubMed]
42. Sly, L.M.; Lopez, M.; Nauseef, W.M.; Reiner, N.E. 1alpha,25-Dihydroxyvitamin D3 -induced monocyte
antimycobacterial activity is regulated by phosphatidylinositol 3-kinase and mediated by the NADPH-
dependent phagocyte oxidase. J. Biol. Chem. 2001, 276, 35482–35493. [CrossRef] [PubMed]
43. Shin, D.-M.; Yuk, J.-M.; Lee, H.-M.; Lee, S.-H.; Son, J.W.; Harding, C.V.; Kim, J.-M.; Modlin, R.L.; Jo, E.-K.
Mycobacterial lipoprotein activates autophagy via TLR2/1/CD14 and a functional vitamin D receptor
signaling. Cell. Microbiol. 2010, 12, 1648–1665. [CrossRef] [PubMed]
44. Yin, Z.; Pintea, V.; Lin, Y.; Hammock, B.D.; Watsky, M.A. Vitamin D enhances corneal epithelial barrier
function. Invest. Ophthalmol. Vis. Sci. 2011, 52, 7359–7364. [CrossRef] [PubMed]
45. Pálmer, H.G.; González-Sancho, J.M.; Espada, J.; Berciano, M.T.; Puig, I.; Baulida, J.; Quintanilla, M.; Cano, A.;
de Herreros, A.G.; Lafarga, M.; et al. Vitamin D3 promotes the differentiation of colon carcinoma cells by
the induction of E-cadherin and the inhibition of beta-catenin signaling. J. Cell Biol. 2001, 154, 369–387.
[CrossRef] [PubMed]
46. D’Amelio, P.; Sassi, F. Gut Microbiota, Immune System, and Bone. Calcif. Tissue Int. 2018, 102, 415–425.
[CrossRef] [PubMed]
47. Caricilli, A.M.; Picardi, P.K.; de Abreu, L.L.; Ueno, M.; Prada, P.O.; Ropelle, E.R.; Hirabara, S.M.; Castoldi, Â.;
Vieira, P.; Camara, N.O.S.; et al. Gut microbiota is a key modulator of insulin resistance in TLR 2 knockout
mice. PLoS Biol. 2011, 9. [CrossRef] [PubMed]
48. Flanagan, P.K.; Chiewchengchol, D.; Wright, H.L.; Edwards, S.W.; Alswied, A.; Satsangi, J.; Subramanian, S.;
Rhodes, J.M.; Campbell, B.J. Killing of escherichia coli by Crohn’s disease monocyte-derived macrophages
and its enhancement by hydroxychloroquine and vitamin D. Inflamm. Bowel Dis. 2015, 21, 1499–1510.
[CrossRef] [PubMed]
49. Su, D.; Nie, Y.; Zhu, A.; Chen, Z.; Wu, P.; Zhang, L.; Luo, M.; Sun, Q.; Cai, L.; Lai, Y.; et al. Vitamin D signaling
through induction of paneth cell defensins maintains gut microbiota and improves metabolic disorders and
hepatic steatosis in animal models. Front Physiol. 2016, 7. [CrossRef] [PubMed]
50. Wu, S.; Liao, A.P.; Xia, Y.; Li, Y.C.; Li, J.-D.; Sartor, R.B.; Sun, J. Vitamin D receptor negatively regulates
bacterial-stimulated NF-kappaB activity in intestine. Am. J. Pathol. 2010, 177, 686–697. [CrossRef] [PubMed]
51. Liu, W.; Chen, Y.; Golan, M.A.; Annunziata, M.L.; Du, J.; Dougherty, U.; Kong, J.; Musch, M.; Huang, Y.;
Pekow, J.; et al. Intestinal epithelial vitamin D receptor signaling inhibits experimental colitis. JCI 2013, 123,
3983–3996. [CrossRef] [PubMed]
87
Nutrients 2018, 10, 1656
52. Golan, M.A.; Liu, W.; Shi, Y.; Chen, L.; Wang, J.; Liu, T.; Li, Y.C. Transgenic expression of vitamin D receptor
in gut epithelial cells ameliorates spontaneous colitis caused by interleukin-10 deficiency. Dig. Dis. Sci. 2015,
60, 1941–1947. [CrossRef] [PubMed]
53. Du, J.; Chen, Y.; Shi, Y.; Liu, T.; Cao, Y.; Tang, Y.; Ge, X.; Nie, H.; Zheng, C.; Li, Y.C. 1,25-Dihydroxyvitamin
D protects intestinal epithelial barrier by regulating the myosin light chain kinase signaling pathway.
Inflamm. Bowel Dis. 2015, 21, 2495–2506. [CrossRef] [PubMed]
54. He, L.; Liu, T.; Shi, Y.; Tian, F.; Hu, H.; Deb, D.K.; Chen, Y.; Bissonnette, M.; Li, Y.C. Gut epithelial vitamin
D receptor regulates microbiota-dependent mucosal inflammation by suppressing intestinal epithelial cell
apoptosis. Endocrinology 2018, 159, 967–979. [CrossRef] [PubMed]
55. Loftus, E.V.; Sandborn, W.J. Epidemiology of inflammatory bowel disease. Gastroenterol. Clin. North Am.
2002, 31, 1–20. [CrossRef]
56. Loftus, E.V. Clinical epidemiology of inflammatory bowel disease: Incidence, prevalence, and environmental
influences. Gastroenterology 2004, 126, 1504–1517. [CrossRef] [PubMed]
57. Lim, W.-C.; Hanauer, S.B.; Li, Y.C. Mechanisms of disease: Vitamin D and inflammatory bowel disease.
Nat. Clin. Pract. Gastroenterol. Hepatol. 2005, 2, 308–315. [CrossRef] [PubMed]
58. Ananthakrishnan, A.N.; Cagan, A.; Gainer, V.S.; Cheng, S.-C.; Cai, T.; Szolovits, P.; Shaw, S.Y.; Churchill, S.;
Karlson, E.W.; Murphy, S.N.; et al. Higher plasma vitamin D is associated with reduced risk of Clostridium
difficile infection in patients with inflammatory bowel diseases. Aliment. Pharmacol. Ther. 2014, 39, 1136–1142.
[CrossRef] [PubMed]
59. Schäffler, H.; Herlemann, D.P.; Klinitzke, P.; Berlin, P.; Kreikemeyer, B.; Jaster, R.; Lamprecht, G. Vitamin D
administration leads to a shift of the intestinal bacterial composition in Crohn’s disease patients, but not in
healthy controls. J. Dig. Dis. 2018, 19, 225–234. [CrossRef] [PubMed]
60. Garg, M.; Hendy, P.; Ding, J.N.; Shaw, S.; Hold, G.; Hart, A. The effect of vitamin D on intestinal inflammation
and faecal microbiota in patients with ulcerative colitis. J. Crohns. Colitis. 2018, 12, 963–972. [CrossRef]
[PubMed]
61. Luthold, R.V.; Fernandes, G.R.; Franco-de-Moraes, A.C.; Folchetti, L.G.D.; Ferreira, S.R.G. Gut microbiota
interactions with the immunomodulatory role of vitamin D in normal individuals. Metab. Clin. Exp. 2017,
69, 76–86. [CrossRef] [PubMed]
62. Bashir, M.; Prietl, B.; Tauschmann, M.; Mautner, S.I.; Kump, P.K.; Treiber, G.; Wurm, P.; Gorkiewicz, G.;
Högenauer, C.; Pieber, T.R. Effects of high doses of vitamin D3 on mucosa-associated gut microbiome vary
between regions of the human gastrointestinal tract. Eur. J. Nutr. 2016, 55, 1479–1489. [CrossRef] [PubMed]
63. Kanhere, M.; He, J.; Chassaing, B.; Ziegler, T.R.; Alvarez, J.A.; Ivie, E.A.; Hao, L.; Hanfelt, J.; Gewirtz, A.T.;
Tangpricha, V. Bolus weekly vitamin D3 supplementation impacts gut and airway microbiota in adults with
cystic fibrosis: A double-blind, randomized, placebo-controlled clinical trial. J. Clin. Endocrinol. Metab. 2018,
103, 564–574. [CrossRef] [PubMed]
64. Bora, S.A.; Kennett, M.J.; Smith, P.B.; Patterson, A.D.; Cantorna, M.T. The gut microbiota regulates endocrine
vitamin D metabolism through fibroblast growth factor 23. Front Immunol. 2018, 9. [CrossRef] [PubMed]
65. Chun, R.F.; Liu, P.T.; Modlin, R.L.; Adams, J.S.; Hewison, M. Impact of vitamin D on immune function:
Lessons learned from genome-wide analysis. Front Physiol. 2014, 5. [CrossRef] [PubMed]
66. Carvalho, J.T.G.; Schneider, M.; Cuppari, L.; Grabulosa, C.C.; T Aoike, D.Q.; Redublo, B.M.C.; Batista, M.;
Cendoroglo, M.; Maria Moyses, R.; Dalboni, M.A. Cholecalciferol decreases inflammation and improves
vitamin D regulatory enzymes in lymphocytes in the uremic environment: A randomized controlled pilot
trial. PLoS ONE 2017, 12. [CrossRef] [PubMed]
67. Xie, Z.; Chen, J.; Zheng, C.; Wu, J.; Cheng, Y.; Zhu, S.; Lin, C.; Cao, Q.; Zhu, J.; Jin, T. 1,25-dihydroxyvitamin
D3 -induced dendritic cells suppress experimental autoimmune encephalomyelitis by increasing proportions
of the regulatory lymphocytes and reducing T helper type 1 and type 17 cells. Immunology 2017, 152, 414–424.
[CrossRef] [PubMed]
68. Zhou, S.-H.; Wang, X.; Fan, M.-Y.; Li, H.-L.; Bian, F.; Huang, T.; Fang, H.-Y. Influence of vitamin D deficiency
on T cell subsets and related indices during spinal tuberculosis. Exp. Ther. Med. 2018, 16, 718–722. [CrossRef]
[PubMed]
69. Stubbs, J.R.; Idiculla, A.; Slusser, J.; Menard, R.; Quarles, L.D. Cholecalciferol supplementation alters
calcitriol-responsive monocyte proteins and decreases inflammatory cytokines in ESRD. J. Am. Soc. Nephrol.
2010, 21, 353–361. [CrossRef] [PubMed]
88
Nutrients 2018, 10, 1656
70. Drozdenko, G.; Heine, G.; Worm, M. Oral vitamin D increases the frequencies of CD38+ human B cells and
ameliorates IL-17-producing T cells. Exp. Dermatol. 2014, 23, 107–112. [CrossRef] [PubMed]
71. Bendix-Struve, M.; Bartels, L.E.; Agnholt, J.; Dige, A.; Jørgensen, S.P.; Dahlerup, J.F. Vitamin D3
treatment of Crohn’s disease patients increases stimulated T cell IL-6 production and proliferation.
Aliment. Pharmacol. Ther. 2010, 32, 1364–1372. [CrossRef] [PubMed]
72. Seibert, E.; Heine, G.H.; Ulrich, C.; Seiler, S.; Köhler, H.; Girndt, M. Influence of cholecalciferol
supplementation in hemodialysis patients on monocyte subsets: A randomized, double-blind,
placebo-controlled clinical trial. Nephron. Clin. Pract. 2013, 123, 209–219. [CrossRef] [PubMed]
73. Buondonno, I.; Rovera, G.; Sassi, F.; Rigoni, M.M.; Lomater, C.; Parisi, S.; Pellerito, R.; Isaia, G.C.;
D’Amelio, P. Vitamin D and immunomodulation in early rheumatoid arthritis: A randomized double-blind
placebo-controlled study. PLoS ONE 2017, 12. [CrossRef] [PubMed]
74. Boonstra, A.; Barrat, F.J.; Crain, C.; Heath, V.L.; Savelkoul, H.F.; O’Garra, A. 1alpha,25-Dihydroxyvitamin
D3 has a direct effect on naive CD4+ T cells to enhance the development of Th2 cells. J. Immunol. 2001, 167,
4974–4980. [CrossRef] [PubMed]
75. Pichler, J.; Gerstmayr, M.; Szépfalusi, Z.; Urbanek, R.; Peterlik, M.; Willheim, M. 1 alpha,25(OH)2 D3 inhibits
not only Th1 but also Th2 differentiation in human cord blood T cells. Pediatr. Res. 2002, 52, 12–18. [PubMed]
76. Staeva-Vieira, T.P.; Freedman, L.P. 1,25-dihydroxyvitamin D3 inhibits IFN-gamma and IL-4 levels during
in vitro polarization of primary murine CD4+ T cells. J. Immunol. 2002, 168, 1181–1189. [CrossRef] [PubMed]
77. Fawaz, L.; Mrad, M.F.; Kazan, J.M.; Sayegh, S.; Akika, R.; Khoury, S.J. Comparative effect of 25(OH)D3 and
1,25(OH)2 D3 on Th17 cell differentiation. Clin. Immunol. 2016, 166, 59–71. [CrossRef] [PubMed]
78. Şıklar, Z.; Karataş, D.; Doğu, F.; Hacıhamdioğlu, B.; İkincioğulları, A.; Berberoğlu, M. Regulatory T cells and
vitamin D status in children with chronic autoimmune thyroiditis. J. Clin. Res. Pediatr. Endocrinol. 2016, 8,
276–281. [CrossRef] [PubMed]
79. Korf, H.; Wenes, M.; Stijlemans, B.; Takiishi, T.; Robert, S.; Miani, M.; Eizirik, D.L.; Gysemans, C.; Mathieu, C.
1,25-Dihydroxyvitamin D3 curtails the inflammatory and T cell stimulatory capacity of macrophages through
an IL-10-dependent mechanism. Immunobiology 2012, 217, 1292–1300. [CrossRef] [PubMed]
80. Mann, E.H.; Ho, T.-R.; Pfeffer, P.E.; Matthews, N.C.; Chevretton, E.; Mudway, I.; Kelly, F.J.; Hawrylowicz, C.M.
Vitamin D counteracts an IL-23-dependent IL-17A+IFN-γ+ response driven by urban particulate matter.
Am. J. Respir. Cell Mol. Biol. 2017, 57, 355–366. [CrossRef] [PubMed]
81. Joshi, S.; Pantalena, L.-C.; Liu, X.K.; Gaffen, S.L.; Liu, H.; Rohowsky-Kochan, C.; Ichiyama, K.; Yoshimura, A.;
Steinman, L.; Christakos, S.; et al. 1,25-dihydroxyvitamin D3 ameliorates Th17 autoimmunity via
transcriptional modulation of interleukin-17A. Mol. Cell. Biol. 2011, 31, 3653–3669. [CrossRef] [PubMed]
82. Chang, J.-H.; Cha, H.-R.; Lee, D.-S.; Seo, K.Y.; Kweon, M.-N. 1,25-Dihydroxyvitamin D3 inhibits the
differentiation and migration of TH17 cells to protect against experimental autoimmune encephalomyelitis.
PLoS ONE 2010, 5. [CrossRef]
83. Colin, E.M.; Asmawidjaja, P.S.; van Hamburg, J.P.; Mus, A.M.C.; van Driel, M.; Hazes, J.M.W.; van
Leeuwen, J.P.T.M.; Lubberts, E. 1,25-dihydroxyvitamin D3 modulates Th17 polarization and interleukin-22
expression by memory T cells from patients with early rheumatoid arthritis. Arthritis Rheum. 2010, 62,
132–142. [CrossRef] [PubMed]
84. Palmer, M.T.; Lee, Y.K.; Maynard, C.L.; Oliver, J.R.; Bikle, D.D.; Jetten, A.M.; Weaver, C.T. Lineage-specific
effects of 1,25-dihydroxyvitamin D3 on the development of effector CD4 T cells. J. Biol. Chem. 2011, 286,
997–1004. [CrossRef] [PubMed]
85. Treiber, G.; Prietl, B.; Fröhlich-Reiterer, E.; Lechner, E.; Ribitsch, A.; Fritsch, M.; Rami-Merhar, B.;
Steigleder-Schweiger, C.; Graninger, W.; Borkenstein, M.; et al. Cholecalciferol supplementation improves
suppressive capacity of regulatory T-cells in young patients with new-onset type 1 diabetes mellitus—A
randomized clinical trial. Clin. Immunol. 2015, 161, 217–224. [CrossRef] [PubMed]
86. Hu, J.; Wan, Y. Tolerogenic dendritic cells and their potential applications. Immunology 2011, 132, 307–314.
[CrossRef] [PubMed]
87. Penna, G.; Adorini, L. 1 Alpha,25-dihydroxyvitamin D3 inhibits differentiation, maturation, activation,
and survival of dendritic cells leading to impaired alloreactive T cell activation. J. Immunol. 2000, 164,
2405–2411. [CrossRef] [PubMed]
89
Nutrients 2018, 10, 1656
88. Piemonti, L.; Monti, P.; Sironi, M.; Fraticelli, P.; Leone, B.E.; Dal Cin, E.; Allavena, P.; Di Carlo, V. Vitamin
D3 affects differentiation, maturation, and function of human monocyte-derived dendritic cells. J. Immunol.
2000, 164, 4443–4451. [CrossRef] [PubMed]
89. Unger, W.W.J.; Laban, S.; Kleijwegt, F.S.; van der Slik, A.R.; Roep, B.O. Induction of Treg by monocyte-derived
DC modulated by vitamin D3 or dexamethasone: Differential role for PD-L1. Eur. J. Immunol. 2009, 39,
3147–3159. [CrossRef] [PubMed]
90. Van Halteren, A.G.S.; Tysma, O.M.; van Etten, E.; Mathieu, C.; Roep, B.O. 1alpha,25-dihydroxyvitamin D3 or
analogue treated dendritic cells modulate human autoreactive T cells via the selective induction of apoptosis.
J. Autoimmun. 2004, 23, 233–239. [CrossRef] [PubMed]
91. Lemire, J.M.; Archer, D.C. 1,25-dihydroxyvitamin D3 prevents the in vivo induction of murine experimental
autoimmune encephalomyelitis. JCI 1991, 87, 1103–1107. [CrossRef] [PubMed]
92. Cantorna, M.T.; Hayes, C.E.; DeLuca, H.F. 1,25-Dihydroxyvitamin D3 reversibly blocks the progression of
relapsing encephalomyelitis, a model of multiple sclerosis. PNAS 1996, 93, 7861–7864. [CrossRef] [PubMed]
93. Cantorna, M.T.; Hayes, C.E.; DeLuca, H.F. 1,25-Dihydroxycholecalciferol inhibits the progression of arthritis
in murine models of human arthritis. J. Nutr. 1998, 128, 68–72. [CrossRef] [PubMed]
94. Alhassan Mohammed, H.; Mirshafiey, A.; Vahedi, H.; Hemmasi, G.; Moussavi NaslKhameneh, A.;
Parastouei, K.; Saboor-Yaraghi, A.A. Immunoregulation of inflammatory and inhibitory cytokines by vitamin
D3 in patients with inflammatory bowel diseases. Scand. J. Immunol. 2017, 85, 386–394. [CrossRef] [PubMed]
95. Tao, C.; Simpson, S.; van der Mei, I.; Blizzard, L.; Havrdova, E.; Horakova, D.; Shaygannejad, V.; Lugaresi, A.;
Izquierdo, G.; Trojano, M.; et al. Higher latitude is significantly associated with an earlier age of disease
onset in multiple sclerosis. J. Neurol. Neurosurg. Psychiatry 2016, 87, 1343–1349. [CrossRef] [PubMed]
96. Laursen, J.H.; Søndergaard, H.B.; Sørensen, P.S.; Sellebjerg, F.; Oturai, A.B. Vitamin D supplementation
reduces relapse rate in relapsing-remitting multiple sclerosis patients treated with natalizumab. Mult. Scler.
Relat. Disord. 2016, 10, 169–173. [CrossRef] [PubMed]
97. Jelinek, G.A.; Marck, C.H.; Weiland, T.J.; Pereira, N.; van der Meer, D.M.; Hadgkiss, E.J. Latitude, sun
exposure and vitamin D supplementation: Associations with quality of life and disease outcomes in a large
international cohort of people with multiple sclerosis. BMC Neurol. 2015, 15. [CrossRef] [PubMed]
98. Mohr, S.B.; Garland, C.F.; Gorham, E.D.; Garland, F.C. The association between ultraviolet B irradiance,
vitamin D status and incidence rates of type 1 diabetes in 51 regions worldwide. Diabetologia 2008, 51,
1391–1398. [CrossRef] [PubMed]
99. Rosecrans, R.; Dohnal, J.C. Seasonal vitamin D changes and the impact on health risk assessment.
Clin. Biochem. 2014, 47, 670–672. [CrossRef] [PubMed]
100. Van der Rhee, H.J.; de Vries, E.; Coebergh, J.W. Regular sun exposure benefits health. Med. Hypotheses 2016,
97, 34–37. [CrossRef] [PubMed]
101. Szilagyi, A.; Leighton, H.; Burstein, B.; Xue, X. Latitude, sunshine, and human lactase phenotype distributions
may contribute to geographic patterns of modern disease: The inflammatory bowel disease model.
Clin. Epidemiol. 2014, 6, 183–198. [CrossRef] [PubMed]
102. Dobson, R.; Giovannoni, G.; Ramagopalan, S. The month of birth effect in multiple sclerosis: Systematic
review, meta-analysis and effect of latitude. J. Neurol. Neurosurg. Psychiatry 2013, 84, 427–432. [CrossRef]
[PubMed]
103. Song, G.G.; Bae, S.-C.; Lee, Y.H. Association between vitamin D intake and the risk of rheumatoid arthritis:
A meta-analysis. Clin. Rheumatol. 2012, 31, 1733–1739. [CrossRef] [PubMed]
104. Zipitis, C.S.; Akobeng, A.K. Vitamin D supplementation in early childhood and risk of type 1 diabetes:
A systematic review and meta-analysis. Arch. Dis. Child. 2008, 93, 512–517. [CrossRef] [PubMed]
105. Dong, J.-Y.; Zhang, W.-G.; Chen, J.J.; Zhang, Z.-L.; Han, S.-F.; Qin, L.-Q. Vitamin D intake and risk of type 1
diabetes: A meta-analysis of observational studies. Nutrients 2013, 5, 3551–3562. [CrossRef] [PubMed]
106. Shen, L.; Zhuang, Q.-S.; Ji, H.-F. Assessment of vitamin D levels in type 1 and type 2 diabetes patients:
Results from metaanalysis. Mol. Nutr. Food Res. 2016, 60, 1059–1067. [CrossRef] [PubMed]
107. Duan, S.; Lv, Z.; Fan, X.; Wang, L.; Han, F.; Wang, H.; Bi, S. Vitamin D status and the risk of multiple sclerosis:
A systematic review and meta-analysis. Neurosci. Lett. 2014, 570, 108–113. [CrossRef] [PubMed]
108. Lin, J.; Liu, J.; Davies, M.L.; Chen, W. Serum vitamin D Level and rheumatoid arthritis disease activity:
Review and meta-analysis. PLoS ONE 2016, 11. [CrossRef] [PubMed]
90
Nutrients 2018, 10, 1656
109. Del Pinto, R.; Pietropaoli, D.; Chandar, A.K.; Ferri, C.; Cominelli, F. Association between inflammatory bowel
disease and vitamin D deficiency: A systematic review and meta-analysis. Inflamm. Bowel Dis. 2015, 21,
2708–2717. [CrossRef] [PubMed]
110. Lu, C.; Yang, J.; Yu, W.; Li, D.; Xiang, Z.; Lin, Y.; Yu, C. Association between 25(OH)D level, ultraviolet
exposure, geographical location, and inflammatory bowel disease activity: A systematic review and
meta-analysis. PLoS ONE 2015, 10. [CrossRef] [PubMed]
111. Sadeghian, M.; Saneei, P.; Siassi, F.; Esmaillzadeh, A. Vitamin D status in relation to Crohn’s disease:
Meta-analysis of observational studies. Nutrition 2016, 32, 505–514. [CrossRef] [PubMed]
112. Feng, R.; Li, Y.; Li, G.; Li, Z.; Zhang, Y.; Li, Q.; Sun, C. Lower serum 25 (OH) D concentrations in type 1
diabetes: A meta-analysis. Diabetes Res. Clin. Pract. 2015, 108, e71–e75. [CrossRef] [PubMed]
113. Pitocco, D.; Crinò, A.; Di Stasio, E.; Manfrini, S.; Guglielmi, C.; Spera, S.; Anguissola, G.B.; Visalli, N.;
Suraci, C.; Matteoli, M.C.; et al. The effects of calcitriol and nicotinamide on residual pancreatic beta-cell
function in patients with recent-onset Type 1 diabetes (IMDIAB XI). Diabet. Med. 2006, 23, 920–923. [CrossRef]
[PubMed]
114. Walter, M.; Kaupper, T.; Adler, K.; Foersch, J.; Bonifacio, E.; Ziegler, A.-G. No effect of the 1alpha,25-
dihydroxyvitamin D3 on beta-cell residual function and insulin requirement in adults with new-onset type 1
diabetes. Diabetes Care 2010, 33, 1443–1448. [CrossRef] [PubMed]
115. Papadimitriou, D.T.; Marakaki, C.; Fretzayas, A.; Nicolaidou, P.; Papadimitriou, A. Negativation of type 1
diabetes-associated autoantibodies to glutamic acid decarboxylase and insulin in children treated with oral
calcitriol. J. Diabetes 2013, 5, 344–348. [CrossRef] [PubMed]
116. Gunasekar, P.; Swier, V.J.; Fleegel, J.P.; Boosani, C.S.; Radwan, M.M.; Agrawal, D.K. Vitamin D and
macrophage polarization in epicardial adipose tissue of atherosclerotic swine. PLoS ONE 2018, 13. [CrossRef]
[PubMed]
117. Yin, K.; You, Y.; Swier, V.; Tang, L.; Radwan, M.M.; Pandya, A.N.; Agrawal, D.K. Vitamin D protects against
atherosclerosis via regulation of cholesterol efflux and macrophage polarization in hypercholesterolemic
swine. Arterioscler. Thromb. Vasc. Biol. 2015, 35, 2432–2442. [CrossRef] [PubMed]
118. Dozio, E.; Briganti, S.; Vianello, E.; Dogliotti, G.; Barassi, A.; Malavazos, A.E.; Ermetici, F.; Morricone, L.;
Sigruener, A.; Schmitz, G.; et al. Epicardial adipose tissue inflammation is related to vitamin D deficiency
in patients affected by coronary artery disease. Nutr. Metab. Cardiovasc. Dis. 2015, 25, 267–273. [CrossRef]
[PubMed]
119. Apostolakis, M.; Armeni, E.; Bakas, P.; Lambrinoudaki, I. Vitamin D and cardiovascular disease. Maturitas
2018, 115, 1–22. [CrossRef] [PubMed]
© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
91
nutrients
Review
The Immunomodulatory and Anti-Inflammatory Role
of Polyphenols
Nour Yahfoufi 1 , Nawal Alsadi 1 , Majed Jambi 1 and Chantal Matar 1,2, *
1 Cellular and Molecular Medicine Department, Faculty of Medicine, University of Ottawa,
Ottawa, ON K1H8L1, Canada; nyahf074@uottawa.ca (N.Y.); nalsa068@uottawa.ca (N.A.);
mjamb055@uottawa.ca (M.J.)
2 School of Nutrition, Faculty of Health Sciences, University of Ottawa, Ottawa, ON K1H8L1, Canada
* Correspondence: Chantal.matar@uottawa.ca; Tel.: +1-613-562-5406
Abstract: This review offers a systematic understanding about how polyphenols target multiple
inflammatory components and lead to anti-inflammatory mechanisms. It provides a clear
understanding of the molecular mechanisms of action of phenolic compounds. Polyphenols regulate
immunity by interfering with immune cell regulation, proinflammatory cytokines’ synthesis, and
gene expression. They inactivate NF-κB (nuclear factor kappa-light-chain-enhancer of activated
B cells) and modulate mitogen-activated protein Kinase (MAPk) and arachidonic acids pathways.
Polyphenolic compounds inhibit phosphatidylinositide 3-kinases/protein kinase B (PI3K/AkT),
inhibitor of kappa kinase/c-Jun amino-terminal kinases (IKK/JNK), mammalian target of rapamycin
complex 1 (mTORC1) which is a protein complex that controls protein synthesis, and JAK/STAT.
They can suppress toll-like receptor (TLR) and pro-inflammatory genes’ expression. Their antioxidant
activity and ability to inhibit enzymes involved in the production of eicosanoids contribute as well to
their anti-inflammation properties. They inhibit certain enzymes involved in reactive oxygen species
ROS production like xanthine oxidase and NADPH oxidase (NOX) while they upregulate other
endogenous antioxidant enzymes like superoxide dismutase (SOD), catalase, and glutathione (GSH)
peroxidase (Px). Furthermore, they inhibit phospholipase A2 (PLA2), cyclooxygenase (COX) and
lipoxygenase (LOX) leading to a reduction in the production of prostaglandins (PGs) and leukotrienes
(LTs) and inflammation antagonism. The effects of these biologically active compounds on the
immune system are associated with extended health benefits for different chronic inflammatory
diseases. Studies of plant extracts and compounds show that polyphenols can play a beneficial role in
the prevention and the progress of chronic diseases related to inflammation such as diabetes, obesity,
neurodegeneration, cancers, and cardiovascular diseases, among other conditions.
1. Introduction
Numerous studies have attributed to polyphenols a broad range of biological activities including
but not limited to anti-inflammatory, immune-modulatory, antioxidant, cardiovascular protective
and anti-cancer actions [1–5]. Polyphenols are ubiquitously made by plants and are present either as
glycosides esters or as free aglycones [6]. More than 8000 structural variants exist in the polyphenol
family. Polyphenols are bioactive compounds found in fruits and vegetables contributing to their color,
flavor, and pharmacological activities [1]. They are classified according to their chemical structures
into flavonoids such as flavones, flavonols, isoflavones, neoflavonoids, chalcones, anthocyanidins, and
proanthocyanidins and nonflavonoids, such as phenolic acids, stilbenoids, and phenolic amides [7].
The majority of these molecules are metabolites of plants, they are made of several aromatic rings
with hydroxyl moieties [8]. Their chemical structures contribute to their classification into different
classes. Considering gastrointestinal digestion, some—but not all—polyphenols are absorbed in
the small intestine, for example, anthocyanins and the majority of remaining polyphenols except
flavonoids are usually stable; these later are unstable in the duodenum. Unabsorbed polyphenols must
be hydrolyzed first by digestive enzymes then glycosides with high lipid contents are absorbed by
epithelial cells [9,10].
In recent years, consumers prefer using natural food ingredients as additives because of their
safety and availability. Applications of phenolic compounds to multiple fresh perishable foods show
that they are worthy to be used as preservatives in foods and can be creditable alternatives to synthetic
food additives. In this sense, polyphenolic compounds start to substitute chemical additives in food.
Different methods like spraying, coating and dipping treatment of food are currently applied in food
technology preceding packaging as effective alternatives [11]. Grape seeds and olive oil polyphenols’
rich extracts can be used as food additives for their anti-oxidant properties [12]. Various polyphenols
like grape polyphenols demonstrate an efficient role as additives in fish and fish products for their
anti-oxidant properties in order to prevent lipid oxidation and quality deterioration of polyunsaturated
fatty acids [13]. In addition polyphenolic compounds like flavonols, p-coumaric, and caffeic acids can
be used as food preservatives for their antimicrobial activity [11].
Back to inflammation, continuous inflammation is known to be a major cause linked to
different human disorders involving cancer, diabetes type II, obesity, arthritis, neurodegenerative
diseases, and cardiovascular diseases [14,15]. Polyphenols derived from botanic origin have shown
anti-inflammatory activity in vitro and in vivo highlighting their beneficial role as therapeutic tools in
multiple acute and chronic disorders [16–20]. Accordingly, many epidemiological and experimental
researches have been studying the anti-inflammatory and immune modulation activities of dietary
polyphenols [15,21]. The ability of these natural compounds to modify the expression of several
pro-inflammatory genes like multiple cytokines, lipoxygenase, nitric oxide synthases cyclooxygenase,
in addition to their anti-oxidant characteristics such as ROS (reactive oxygen species) scavenging
contributes to the regulation of inflammatory signaling [22,23]. This review will discuss the
immunomodulatory effects of dietary polyphenols, their anti-inflammatory abilities, the different
mechanisms and pathways involved in reducing inflammation and their contribution to protect from
different chronic inflammatory diseases with a focus on their anti-cancer activity.
93
Nutrients 2018, 10, 1618
human endothelial cells MAPK (mitogen-activated protein Kinase), and IKK (inhibitor of kappa
kinase). Moreover, curcumin downregulates NF-κB (nuclear factor kappa-light-chain-enhancer of
activated B cells) and STAT3 (signal transducer and activator of transcription) and reduces the
expression of TLR-2 (toll-like receptor-2) and 4 while, in vivo, it upregulates PPARγ (Peroxisome
proliferator-activated receptor gamma) in male adult rats [30–35]. Caffeic acid phenethyl ester
suppresses TLR4 activation and LPS-mediated NF-κB in macrophages, Quercetin was also shown to
inhibit leukotriens biosynthesis in human polymorphonuclear leukocytes [36,37]. COX2 expression is
also attenuated by ECGC (Epigallocatechin gallate) in colon cancer cell and androgen-independent
PC-3 cells of human prostate cancer, gingerol in and piceatannol (EGCG analog found in Norway
spruces) leading to NFκ B inactivation [30,38–40]. Furthermore, polyphenols, such as Gingerol and
Quercetin can activate the production of adiponectin known for its anti-inflammatory effects [30,39].
Similarily, EGCG blocks NFκ B activation in human epithelial cells and downregulates the expression
of iNOS (inducible nitric oxide synthase), NO (nitric oxide) production in macrophages resulting in its
immunomodulation [38,40,41]. A series of in vitro studies found that other polyphenols like oleanolic
acid, curcumin, kaempferol-3-O-sophoroside, EGCG and lycopene inhibit high mobility group box1
protein, an important chromatin protein that interacts with nucleosomes, transcription factors, and
histones regulating transcription and playing a key role in inflammation [35]. All of these examples
support the anti-inflammatory effects of polyphenols.
Polyphenols’ use is associated with a direct change in the count and differentiation of specific
immune cells. An increase in T helper 1(Th1), natural killer (NK), macrophages and dendritic cells (DCs)
in Peyer’s patches and spleen is associated with oral administration of polyphenols extracted from the
fruit date in male C3H/HeN mice [24]. In humans, the count of regulatory T cells (Treg or suppressor T
cells) characterized by the (CD4 + CD25 + Foxp3+) phenotype and involved in immune tolerance and
autoimmune control can be boosted by polyphenols [42–44]. In vivo, Epigallocatechin-3-gallate, found
in green tea and injected to Laboratory inbred strain (BALB)/c mice, rises the number of functional Treg
in spleens, pancreatic lymph nodes, and mesentheric lymph nodes [45]. Similarly, in vitro treatment of
Jurkat T cells with EGCG or green tea upsurges the expression of Foxp3 and IL10. Baicalin, a flavone,
extracted from Huangqin herb, induces Foxp3 expression in HEK 293 T cells and triggers functional
Treg from splenic CD4 + CD25− T cells [46]. Additionally, flavonoids show an agonistic effect of aryl
hydrocarbon receptor (AhR) and bind xenobiotic-responsive elements in promoter regions of certain
genes, including Foxp3 rising its expression [47].
Th1 and Th17 populations are also affected by polyphenols: EGCG reduces the differentiation of
Th1 and reduces the numbers of Th17 and Th9 cells in specific pathogen-free C57/BL6 female mice [48].
Other polyphenols like Baicalin show a reduction of Th17 differentiation in vitro and a diminution of
IL-17 expression [49].
Macrophages are affected by polyphenols as well. Macrophages are known to be a key
player in the inflammatory response. They initiate inflammation by secreting pro-inflammatory
mediators and cytokines like IL-6 and TNF-α [50]. Polyphenols repress macrophages by inhibiting
cyclooxygenase-2 (COX-2), inducible nitric oxide synthase (iNOS), thus they reduce the production of
TNF-α, interleukine-1-beta (IL-1-β) and IL-6 expression [51]. Chinese propolis [52] containing ferulic
acid and coumaric acid, an extract of Lonicera japónica Thunb) [53] or Kalanchoe gracilis [54] are a good
example in this case as per demonstrated by in vitro studies using RAW 264.7 cells.
94
Nutrients 2018, 10, 1618
and in vitro studies demonstrate that polyphenols affect macrophages by inhibiting multiple key
regulators of inflammatory response such as the inhibition of TNF-α, IL-1-β, and IL-6 [51].
In humans, consumption of bilberries is associated with a decreased inflammation score in
patients’ blood, reflected by decreasing serum levels of IL-6, IL-12, and high sensitivity C reactive
protein [56]. Moreover, clinical trials have shown the ability of polyphenol-enriched extra virgin olive
oil to reduce IL-6 and C-reactive protein expression in stable coronary heart disease patients [57].
In lipopolysaccharide (LPS)-treated BALB/c mice, a model system of inflammation olive
vegetation water show ability to inhibit the production of tumor necrosis factor-alpha usually
activated by inflammation [58]. Flavonoids, as well, play an important anti-inflammatory effect
by influencing cytokines’ secretion. Several flavonoids are found able to inhibit the expression
of various pro-inflammatory cytokines and chemokines like TNFα, IL-1β, IL-6, IL-8, and MCP-1
(monocyte chemoattractant protein-1) in multiple cell types such as LPS-activated mouse primary
macrophages, activated human mast cell line, activated human astrocytes, human synovial cells, and
human peripheral blood mononuclear cells [59–64]. In murine RAW 264.7 macrophages stimulated by
LPS, Chinese propolis as well as extract of Lonicera japónica Thunb (Caprifoliaceae) or Kalanchoe gracilis
demonstrated inhibitory effects on TNF-α, IL-1-β, and IL-6 [52–54]. Similarly, certain polyphenol
analogs, like curcumin analog EF31, have shown the ability to inhibit the expression and secretion of
TNF-α, IL-1-β, and IL-6 in mouse Raw 264.7 macrophages [65].
Likewise, reduction of the secretion of TNF-α and IL-6 without IL-1β modulation is observed
with extracts of chamomile, meadowsweet, willow bark, and isolated polyphenols such as quercetin
existing in these extracts in THP1 macrophages [66]. Extract of Cydonia oblonga inhibits TNF-α and
Interleukin 8 while it increases IL-10 and IL-6 in THP-1monocytes stimulated with LPS. The reduction
in TNF-α levels limits the acute inflammatory response [67,68]. Other cytokines like IFNγ might also
be inhibited by certain polyphenols. For example, kaempferol reduces the production of IFN-γ in a
dose-dependent manner in spleen cells and T cell lines [69].
Certain polyphenols exert their effects on the balance between pro- and anti-inflammatory
cytokines production such as quercetin and catechins, they enhance IL-10 release while they
inhibit TNFα and IL-1β [59,70]. Extract of Cydonia oblonga also inhibits the effects of TNF-α and
Interleukin 8 (IL-8) while it raises IL-10 in the same type of monocytes [67,68]. Modulation of
inflammatory cytokines is one of many common mechanisms by which polyphenols in general
exert their immunomodulatory effects.
95
Nutrients 2018, 10, 1618
of IKK, and the processing of NFκ B precursors [73–75]. Thus, the inhibition of NFκB can be of
a great benefit in controlling inflammatory conditions [76]. Several polyphenols modulate NFκ B
activation and reduce inflammation [77,78]. Quercetin blocks the nuclear translocation of p50 and
p65 subunits of NFκ B and represses the expression of pro-inflammatory associated genes, NOS and
COX-2 in RAW264.7 macrophages [79]. It inhibits the phosphorylation of Iκ Bα protein both in vitro
(using macrophages) and in vivo (using dextran sulfate sodium (DSS) rat colitis model) leading to
inactivation of the NFκ B pathway [80]. In human mast cells, quercetin prevents the degradation
of Iκ Bα, as well as the nuclear translocation of p65 resulting in reduction of TNFα, IL-1β, IL-6 and
IL-8 [63]. It can modulate chromatin remodeling, for example it blocks the recruitment of a histone
acetyl transferase called CBP/p300 to the promoters of interferon-inducible protein 10 (IP-10) and
macrophage inflammatory protein-2 (MIP-2) genes in primary murine small intestinal epithelial cell.
As a result, it inhibits the expression of these pro-inflammatory cytokines [81]. Quercetin can block
the activation of IKK, NFκ B, and it reduces the ability of NFκ B to bind DNA in microglia treated
by LPS and IFN-γ in mouse BV-2 microglia [82]. Luteolin, too, blocks NFκ B activation and inhibits
pro-inflammatory genes expression and the cytokines production in murine macrophages RAW 264.7
and mouse alveolar macrophages; it also inhibits IKKs in LPS-induced epithelial and dendritic cells [83].
In addition, in co-cultured intestinal epithelial Caco-2 and macrophage RAW 264.7 cells, luteolin
represses NF-k B activation and TNF-α secretion [84]. Likewise, Genistein represses LPS-induced
activation of NF-k B in monocytes and reduces the inflammation by inhibiting NF-k B activation
upon adenosine monophosphate activated protein kinase stimulation in LPS-stimulated macrophages
RAW 264.7 [83,85]. Galangin, as well, stops the degradation of Ik Bα and the translocation of p65 NF-k
B, repressing the expression of TNF-α, IL-6, IL-1β, and IL-8 in mast cell [86]. EGCG counteracts the
activation of IKK and the degradation of Iκ Bα and inhibits NFκ B in culture respiratory epithelial
cells and in vivo in male Wistar rats [87,88]. Furthermore, EGCG blocks DNA binding of NFκ B which
reduces the expression of IL-12p40 and iNOS in murine peritoneal macrophages [89,90]. Catechin and
epichatechin reduce NFκ B activity in PMA-induced Jurkat T cells. Flavonoids can modulate NFκ B
activation cascade at early phases by affecting IKK activation and regulation of oxidant levels or at late
phases by affecting binding of NF-κ B to DNA in jurkat Tcells [91]. Hydroxytyrosol, and resveratrol
inhibit NFκ B activation, and the expression of VCAM-1 in LPS-stimulated human umbilical vein
endothelial cells [92]. In summary, polyphenols can modulate NFκ B activation cascade at different
steps such as by affecting IKK activation and regulating of the oxidant levels or by affecting binding of
NF-κ B to DNA leading to an important anti-inflammatory effect responsible for their potential value
in treating chronic inflammatory conditions (Figure 1).
Figure 1. Potential points of action of polyphenols within inflammatory cascade. NF-κ B: nuclear factor
kappa-light-chain-enhancer of activated B cells; IKK: IkB-kinase; ERK: extracellular signal-related
kinases; JNK: c-Jun amino-terminal kinases; p38 (or p38-MAPK): p38-mitogen-activated protein kinase;
COX: cyclooxygenase; LOX: lipoxygenase; AA: arachidonic acid; PLA2: phospholipase A2; PGs:
prostaglandins; LTs: leukotriens. For references see the text.
96
Nutrients 2018, 10, 1618
97
Nutrients 2018, 10, 1618
other polyphenols, such as kaempferol in culture of LPS-stimulated human whole blood cells [117].
Extra virgin olive oil rich with more than 30 phenolic compounds inhibit 5-LOX in human activated
leukocytes reducing leukotriene B4 and suppresses eicosanoids production by animal and human
cells in vitro [118,119]. Finally, certain polyphenols show structural and functional similarities with
specific anti-inflammatory drugs. A phenolic compound—oleocanthal—demonstrates a natural
anti-inflammatory property and exhibits structural similarities to the ibuprofen (a well-known
anti-inflammatory drug). Oleocanthal—like ibuprofen—inhibits COX-1 and COX-2 activities in a
dose-dependent manner [120].
Transition metal ions, like Fe+2 , Cu2+ , Co2+ , Ti3+ , or Cr5+ , results in OH• formation from
H2 O2 [131,132]. Curcumin is able to chelate transition metal (Cu2+ and Fe2+ ) ions. Alike, EGCG
and quercetin chelate Fe2+ (iron ion) [128]. Polyphenols like apocynin, reservatol, and curcumin
can inhibit NOX (NADPH oxidase) causing a reduction in the generation of O2 • during infections
consecutively in endothelial cells in THP1-monocytes [133–135]. Additionally, polyphenols can
attenuate the mitochondrial ATP synthesis by blocking the mitochondrial respiratory chain and
ATPase. As a result, ROS production is diminished. Curcumin [136], EGCG [137], phenolic acids [138],
capsaicin [139], quercetins [140], anthocyanins [140], and resveratrol analogs [141] inhibit xanthine
oxidase. Thus, they reduce ROS production. Polyphenols affect the activity of cyclooxygenase,
98
Nutrients 2018, 10, 1618
lipoxygenase, and NOS (nitric oxide synthase) as per found in RAW 264.7 macrophes [142]. These
enzymes are known to metabolize arachidonic acid and their inhibition moderates the production
of key mediators of inflammation (prostaglandins, leukotrienes, and NO . . . ) [142]. Polyphenols
can also restrain LPS-induced iNOS gene expression in cultured macrophages, decreasing oxidative
harm [143]. Finally, they may act by upregulating endogenous antioxidant enzymes. In vivo, curcumin
can stimulate antioxidant enzymes like superoxide dismutase (SOD), catalase, and glutathione (GSH)
peroxidase (Px) which lead to ROS detoxification [144]. Likewise, EGCG rises SOD and GSH-Px
activities with augmented amount of cellular glutathione [145]. In conclusion, polyphenols exert
the anti-inflammatory action by different mechanisms: Radical scavenging, metal chelating, NOX
inhibition, tempering the mitochondrial respiratory chain, inhibition of certain enzymes involved in
ROS production, like xanthine oxidase and upregulation of endogenous antioxidant enzymes.
99
Nutrients 2018, 10, 1618
100
Nutrients 2018, 10, 1618
phase II enzymes that initiate the formation of polar metabolites which are readily excreted from the
body [195]. Certain dietary polyphenols such as flavonoids reduce cellular formation of ROS which
protects from the oxidation of DNA [193].
In addition to their anti-oxidant properties, pro-oxidant characteristic of polyphenols is important
in treating and preventing cancer. Pro-oxidant activity can be initiated by certain conditions such
as superoxide leakage [196]. The pro-oxidant activities of polyphenols in cancer cells can result in
inducing apoptosis [197], cell cycle arrest [198] and inhibiting the proliferation signaling pathways
(i.e., epidermal growth factor receptor/mitogen activated protein kinase, phosphatidylinositide
3-kinases/protein kinase B, as well as NF-kB) [199]. For example, polyphenols from apple are able
to inhibit the proliferation of human bladder transitional cell carcinoma (TCC, TSGH-8301 cells),
inducing G2/M cell cycle arrest, and promoting apoptosis [200]. In human papilloma virus-18-positive
HeLa cervical cancer cells, green tea polyphenols can induce cell cycle arrest at the subG1 phase and
apoptosis through caspases activation [201]. Flavonoids, such as quercetin, induce apoptosis in many
cancer cells such as leukemic U937 cell [202], prostate cancer cells [203], hepatic cancer cells [204],
among other types. A combination of quercetin with resveratrol and catechin inhibits breast cancer
progression in vitro and in vivo by inducing apoptosis in carcinogenic breast cells [205]. In addition,
polyphenols can reduce cancer metastasis such as quercetin [206,207].
Sufficient studies have reported that NF-κB signaling pathways are closely related to cancer
metastasis. Polyphenols can disrupt the metastatic potential of cancer by inhibiting NF-κB activity [208].
Curcumin is a good example [209–211] of decreasing cancer metastasis in mice by suppressing
NF-κB expression and down-regulating VEGF (vascular endothelial growth factor), COX-2, and
MMP-9 (matrix metallopeptidase-9) expression in tissues of the breast, brain, lung, liver, and
spleen [212,213]. Moreover, the strength of metastasis is associated to the epithelial-to-mesenchymal
transition (EMT) [214]. There is robust evidence that polyphenols compounds can modulate EMT
and its related signaling pathways [215]. For example, EGCG, a flavan-3-ol, induces apoptosis and
significantly reduces colony formation and cell migration in nasopharyngeal carcinoma (NPC) and
cancer stem cells (CSC) in different cell lines [216]. Luteolin and quercetin reverse the migration and
invasiveness of metastatic cells by reducing the expression of mesenchymal markers and transcriptional
factors on the cell membrane (i.e., twist, snail, and N-cadherin) and upregulating adhesion molecules
such as E-cadherin [217]. Thus, through variable mechanisms, polyphenols broadly downregulate
inflammation origination, progression, and evolution to cancers (Figure 3).
101
Nutrients 2018, 10, 1618
7. Conclusions
In conclusion, the vast number of published studies proved the immunomodulatory role of
polyphenols in vivo and in vitro. Different underlying regulatory mechanisms are now well elucidated.
These data highlight the promising role of polyphenols in prevention and therapy of diseases with
underlining inflammatory conditions, including cancer, neurodegenerative diseases, obesity, type
II diabetes, and cardiovascular diseases. However, the role of polyphenols in modulating multiple
inflammatory cellular pathways should be further investigated. Many questions remain unanswered
about the usage of polyphenols in clinical setting. The role of the microbiota in degrading these
polyphenols should be further studied. The notion of bioavailability and its impact on biofunctionality
should also be revisited. It is generally believed that polyphenol activity is principally located in the gut
where their immunoprotective and anti-inflammatory activities are initiated and subsequently ensuring
systemic anti-inflammatory effects. Since different polyphenols can have multiple intracellular
targets, additional data is needed to determine the consequences of the interaction or the synergistic
effects between multiple polyphenolic compounds or polyphenols and commonly used medications.
Moreover, further in vivo and meta-analysis studies in humans are necessary to fully reveal the
mechanisms of action of polyphenols in several physiological conditions in order to produce important
insights into their prophylactic and therapeutic uses.
Author Contributions: N.Y. wrote all parts of the paper and prepared all Figures as well as the graphabstract
except cancer and polyphenol paragraph and Figure 3; N.A. wrote cancer and polyphenol paragraph and prepared
Figure 3; M.J. collected some papers and revised the Figures and their design; C.M. revised and guided the work.
Funding: This research received no external funding.
Acknowledgments: Special thanks to uOttawa libraries especially health sciences library and Morisset library
and the department of “Nutrition Sciences”.
Conflicts of Interest: The authors declare no conflict of interest.
Abbreviations
ROS Reactive oxygen species
COX Cyclooxygenase
NOX NADPH oxidase
SOD Superoxide dismutase
GSH Glutathione
Px Peroxidase
PLA2 Phospholipase A2
PGs Prostaglandins
LTs Leukotrienes
MAPK Mitogen-activated protein Kinase
IKK Inhibitor of kappa kinase
NFκB Nuclear factor kappa-light-chain-enhancer of activated B cells
Th1 T helper 1
NK Natural killer
DCs Dendritic cells
ECGC Epigallocatechin gallate
Treg Regulatory T cells
102
Nutrients 2018, 10, 1618
References
1. Recio, M.; Andujar, I.; Rios, J. Anti-Inflammatory Agents from Plants: Progress and Potential. Curr. Med. Chem.
2012, 19, 2088–2103. [CrossRef] [PubMed]
2. Eberhardt, M.V.; Lee, C.Y.; Liu, R.H. Antioxidant Activity of Fresh Apples. Nature 2000, 405, 903–904.
[CrossRef] [PubMed]
3. Spagnuolo, C.; Russo, M.; Bilotto, S.; Tedesco, I.; Laratta, B.; Russo, G.L. Dietary Polyphenols in Cancer
Prevention: The Example of the Flavonoid Quercetin in Leukemia. Ann. N. Y. Acad. Sci. 2012, 1259, 95–103.
[CrossRef] [PubMed]
4. Andriantsitohaina, R.; Auger, C.; Chataigneau, T.; Étenne-Selloum, N.; Li, H.; Martínez, M.C.;
Schini-Kerth, V.B.; Laher, I. Molecular Mechanisms of the Cardiovascular Protective Effects of Polyphenols.
Br. J. Nutr. 2012, 108, 1532–1549. [CrossRef] [PubMed]
5. Vauzour, D.; Rodriguez-Mateos, A.; Corona, G.; Oruna-Concha, M.J.; Spencer, J.P.E. Polyphenols and Human
Health: Prevention of Disease and Mechanisms of Action. Nutrients 2010, 2, 1106–1131. [CrossRef] [PubMed]
6. Ma, Y.; Kosinska-Cagnazzo, A.; Kerr, W.L.; Amarowicz, R.; Swanson, R.B.; Pegg, R.B. Separation and
Characterization of Soluble Esterified and Glycoside-Bound Phenolic Compounds in Dry-Blanched Peanut
Skins by Liquid Chromatography-Electrospray Ionization Mass Spectrometry. J. Agric. Food Chem. 2014, 62,
11488–11504. [CrossRef] [PubMed]
103
Nutrients 2018, 10, 1618
7. Tsao, R. Chemistry and Biochemistry of Dietary Polyphenols. Nutrients 2010, 2, 1231–1246. [CrossRef]
[PubMed]
8. Cheynier, V. Polyphenols in Food Are More Complex Then Often Thought. Am. J. Clin. Nutr. 2005, 81,
223–229. [CrossRef] [PubMed]
9. Mosele, J.I.; Macia, A.; Romero, M.-P.; Motilua, M.-J.; Rubio, L. Application of in Vitro Gastrointestinal
Digestion and Colonic\nfermentation Models to Pomegranate Products (Juice, Pulp and Peel\nextract) to
Study the Stability and Catabolism of Phenolic Compounds. J. Funct. Food 2015, 14, 529–540. [CrossRef]
10. Correa-Betanzo, J.; Allen-Vercoe, E.; McDonald, J.; Schroeter, K.; Corredig, M.; Paliyath, G. Stability and
Biological Activity of Wild Blueberry (Vaccinium Angustifolium) Polyphenols during Simulated in Vitro
Gastrointestinal Digestion. Food Chem. 2014, 165, 522–531. [CrossRef] [PubMed]
11. Martillanes, S.; Rocha-Pimienta, J.; Cabrera-Bañegil, M.; Martín-Vertedor, D.; Delgado-Adámez, J. Application
of Phenolic Compounds for Food Preservation: Food Additive and Active Packaging. In Phenolic
Compounds-Biological Activity; InTech: London, UK, 2017.
12. Maqsood, S.; Benjakul, S.; Shahidi, F. Emerging Role of Phenolic Compounds as Natural Food Additives in
Fish and Fish Products. Crit. Rev. Food Sci. Nutr. 2013, 53, 162–179. [CrossRef] [PubMed]
13. Maestre, R.; Micol, V.; Funes, L.; Medina, I. Incorporation and Interaction of Grape Seed Extract in Membranes
and Relation with Efficacy in Muscle Foods. J. Agric. Food Chem. 2010, 58, 8365–8374. [CrossRef] [PubMed]
14. Kennedy, E.T. Evidence for Nutritional Benefits in Prolonging Wellness. Am. J. Clin. Nutr. 2006, 8, 16470004.
[CrossRef] [PubMed]
15. Bengmark, S. Acute and “Chronic” Phase Reaction-a Mother of Disease. Clin. Nutr. 2004, 23, 1256–1266.
[CrossRef] [PubMed]
16. Visioli, F.; Galli, C. The Effect of Minor Constituents of Olive Oil on Cardiovascular Disease: New Findings.
Nutr. Rev. 1998, 56, 142–147. [CrossRef] [PubMed]
17. Visioli, F.; Galli, C. The Role of Antioxidants in the Mediterranean Diet. Lipids 2001, 36, S49–S52. [CrossRef]
[PubMed]
18. Middleton, E., Jr.; Kandaswami, C.; Theoharides, T.C. The Effects of Plant Flavonoids on Mammalian Cells:
Implications for Inflammation, Heart Disease, and Cancer. Pharmacol. Rev. 2000, 52, 673–751. [PubMed]
19. Urquiaga, J.; Leighton, F. Plant Polyphenol Antioxidants and Oxidative Stress. Biol. Res. 2000, 33, 55–64.
[CrossRef] [PubMed]
20. Scalbert, A.; Manach, C.; Morand, C.; Rémésy, C.; Jiménez, L. Dietary Polyphenols and the Prevention of
Diseases. Crit. Rev. Food Sci. Nutr. 2005, 45, 287–306. [CrossRef] [PubMed]
21. Yoon, J.H.; Baek, S.J. Molecular Targets of Dietary Polyphenols with Anti-Inflammatory Properties.
Yonsei Med. J. 2005, 46, 585–596. [CrossRef] [PubMed]
22. Malireddy, S.; Kotha, S.R.; Secor, J.D.; Gurney, T.O.; Abbott, J.L.; Maulik, G.; Maddipati, K.R.; Parinandi, N.L.
Phytochemical Antioxidants Modulate Mammalian Cellular Epigenome: Implications in Health and Disease.
Antioxid. Redox Signal. 2012, 17, 327–339. [CrossRef] [PubMed]
23. Santangelo, C.; Varì, R.; Scazzocchio, B.; Di Benedetto, R.; Filesi, C.; Masella, R. Polyphenols, Intracellular
Signalling and Inflammation. Ann. Ist. Super. Sanita 2007, 43, 394–405. [PubMed]
24. Karasawa, K.; Uzuhashi, Y.; Hirota, M.; Otani, H. A Matured Fruit Extract of Date Palm Tree
(Phoenix dactylifera L.) Stimulates the Cellular Immune System in Mice. J. Agric. Food Chem. 2011, 59,
11287–11293. [CrossRef] [PubMed]
25. John, C.M.; Sandrasaigaran, P.; Tong, C.K.; Adam, A.; Ramasamy, R. Immunomodulatory Activity of
Polyphenols Derived from Cassia Auriculata Flowers in Aged Rats. Cell. Immunol. 2011, 271, 474–479.
[CrossRef] [PubMed]
26. Mohar, D.; Malik, S. The Sirtuin System: The Holy Grail of Resveratrol? J. Clin. Exp. Cardiol. 2012, 3, 216.
[CrossRef] [PubMed]
27. Speciale, A.; Chirafisi, J.; Saija, A.; Cimino, F. Nutritional Antioxidants and Adaptive Cell Responses:
An Update. Curr. Mol. Med. 2011, 11, 770–789. [CrossRef] [PubMed]
28. Biasutto, L.; Mattarei, A.; Zoratti, M. Resveratrol and Health: The Starting Point. ChemBioChem 2012, 13,
1256–1259. [CrossRef] [PubMed]
29. Capiralla, H.; Vingtdeux, V.; Venkatesh, J.; Dreses-werringloer, U.; Zhao, H.; Davies, P.; Marambaud, P.
Identification of Potent Small? Molecule Inhibitors of STAT3 with Anti? Inflammatory Properties in RAW
264.7 Macrophages. FEBS J. 2012, 279, 3791–3799. [CrossRef] [PubMed]
104
Nutrients 2018, 10, 1618
30. Leiherer, A.; Mündlein, A.; Drexel, H. Phytochemicals and Their Impact on Adipose Tissue Inflammation
and Diabetes. Vasc. Pharmacol. 2013, 58, 3–20. [CrossRef] [PubMed]
31. Siddiqui, A.M.; Cui, X.; Wu, R.; Dong, W.; Zhou, M.; Hu, M.; Simms, H.H.; Wang, P. The Anti-Inflammatory
Effect of Curcumin in an Experimental Model of Sepsis Is Mediated by up-Regulation of Peroxisome
Proliferator-Activated Receptor-γ. Crit. Care Med. 2006, 34, 1874–1882. [CrossRef] [PubMed]
32. Marchiani, A.; Rozzo, C.; Fadda, A.; Delogu, G.; Ruzza, P. Curcumin and Curcumin-like Molecules: From
Spice to Drugs. Curr. Med. Chem. 2014, 21, 204–222. [CrossRef] [PubMed]
33. Noorafshan, A.; Ashkani-Esfahani, S. A Review of Therapeutic Effects of Curcumin. Curr. Pharm. Des. 2013,
19, 2032–2046. [PubMed]
34. Gupta, S.C.; Prasad, S.; Kim, J.H.; Patchva, S.; Webb, L.J.; Priyadarsini, I.K.; Aggarwal, B.B. Multitargeting by
Curcumin as Revealed by Molecular Interaction Studies. Nat. Prod. Rep. 2011, 28, 1937–1955. [CrossRef]
[PubMed]
35. Bae, J. Role of High Mobility Group Box 1 in Inflammatory Disease: Focus on Sepsis. Arch. Pharm. Res. 2012,
35, 1511–1523. [CrossRef] [PubMed]
36. Tsuda, S.; Egawa, T.; Ma, X.; Oshima, R.; Kurogi, E.; Hayashi, T. Coffee Polyphenol Caffeic Acid but Not
Chlorogenic Acid Increases 5’AMP-Activated Protein Kinase and Insulin-Independent Glucose Transport in
Rat Skeletal Muscle. J. Nutr. Biochem. 2012, 23, 1403–1409. [CrossRef] [PubMed]
37. Akyol, S.; Ozturk, G.; Ginis, Z.; Amutcu, F.; Yigitoglu, M.; Akyol, O. In Vivo and in Vitro Antıneoplastic
Actions of Caffeic Acid Phenethyl Ester (CAPE): Therapeutic Perspectives. Nutr. Cancer 2013, 65, 1515–1526.
[CrossRef] [PubMed]
38. Kanwar, J. Recent Advances on Tea Polyphenols. Front. Biosci. 2012, E4, 111–131. [CrossRef]
39. Domitrovic, R. The Molecular Basis for the Pharmacological Activity of Anthocyans. Curr. Med. Chem. 2011,
18, 4454–4469. [CrossRef] [PubMed]
40. Singh, B.; Shankar, S.; Sriivastava, R. Green Tea Catechin, Epigallocatechin-3-Gallate (EGCG): Mechanisms,
Perspectives and Clinical. Biochem. Pharmacol. 2011, 82, 1807–1821. [CrossRef] [PubMed]
41. Landis-Piwowar, K.; Chen, D.; Foldes, R.; Chan, T.-H.; Dou, Q.P. Novel Epigallocatechin Gallate Analogs
as Potential Anticancer Agents: A Patent Review (2009–Present). Expert Opin. Ther. Pat. 2013, 23, 189–202.
[CrossRef] [PubMed]
42. Sakaguchi, S.; Miyara, M.; Costantino, C.M.; Hafler, D.A. FOXP3 + Regulatory T Cells in the Human Immune
System. Nat. Rev. Immunol. 2010, 10, 490–500. [CrossRef] [PubMed]
43. Boissier, M.C.; Assier, E.; Biton, J.; Denys, A.; Falgarone, G.; Bessis, N. Regulatory T Cells (Treg) in Rheumatoid
Arthritis. J. Bone Spine 2009, 76, 10–14. [CrossRef] [PubMed]
44. Robinson, D.S.; Larché, M.; Durham, S.R. Tregs and Allergic Disease. J. Clin. Investig. 2004, 114, 1389–1397.
[CrossRef] [PubMed]
45. Wong, C.P.; Nguyen, L.P.; Noh, S.K.; Bray, T.M.; Bruno, R.S.; Ho, E. Induction of Regulatory T Cells by Green
Tea Polyphenol EGCG. Immunol. Lett. 2011, 139, 7–13. [CrossRef] [PubMed]
46. Yang, J.; Yang, X.; Li, M. Baicalin, a Natural Compound, Promotes Regulatory T Cell Differentiation.
IBMC Complement. Altern. Med. 2012, 16, 64. [CrossRef] [PubMed]
47. Wang, H.K.; Yeh, C.H.; Iwamoto, T.; Satsu, H.; Shimizu, M.; Totsuka, M. Dietary Flavonoid Naringenin
Induces Regulatory T Cells via an Aryl Hydrocarbon Receptor Mediated Pathway. J. Agric. Food Chem. 2012,
60, 2171–2178. [CrossRef] [PubMed]
48. Wang, J.; Pae, M.; Meydani, S.N.; Wu, D. Green Tea Epigallocatechin-3-Gallate Modulates Differentiation of
Naïve CD4+T Cells into Specific Lineage Effector Cells. J. Mol. Med. 2013, 91, 485–495. [CrossRef] [PubMed]
49. Yang, J.; Yang, X.; Chu, Y.; Li, M. Identification of Baicalin as an Immunoregulatory Compound by Controlling
TH17 Cell Differentiation. PLoS ONE 2011, 6, e21359178. [CrossRef] [PubMed]
50. Murray, P.J.; Wynn, T.A. Protective and Pathogenic Functions of Macrophage Subsets. Nat. Rev. Immunol.
2011, 11, 723–737. [CrossRef] [PubMed]
51. González, R.; Ballester, I.; López-Posadas, R.; Suárez, M.D.; Zarzuelo, A.; Martínez-Augustin, O.; Sánchez de
Medina, F. Effects of Flavonoids and Other Polyphenols on Inflammation. Crit. Rev. Food Sci. Nutr. 2011, 51,
331–362. [CrossRef] [PubMed]
52. Wang, K.; Ping, S.; Huang, S.; Hu, L.; Xuan, H.; Zhang, C.; Hu, F. Molecular Mechanisms Underlying the In
Vitro Anti-Inflammatory Effects of a Fla Vonoid -Rich Ethanol Extract from Chinese Propolis (Poplar Type).
Cell 2013, 2013, 127672.
105
Nutrients 2018, 10, 1618
53. Park, K.I.; Kang, S.R.; Park, H.S.; Lee, D.H.; Nagappan, A.; Kim, J.A.; Shin, S.C.; Kim, E.H.; Lee, W.S.;
Chung, H.J.; et al. Regulation of Proinflammatory Mediators via NF-KB and P38 MAPK-Dependent
Mechanisms in RAW 264.7 Macrophages by Polyphenol Components Isolated from Korea Lonicera Japonica
THUNB. Evid.-Based Complement. Altern. Med. 2012, 2012, 22611435. [CrossRef] [PubMed]
54. Lai, Z.-R.; Ho, Y.-L.; Huang, S.-C.; Huang, T.-H.; Lai, S.-C.; Tsai, J.-C.; Wang, C.-Y.; Huang, G.-J.; Chang, Y.-S.
Antioxidant, Anti-Inflammatory and Antiproliferative Activities of Kalanchoe gracilis (L.) DC Stem. Am. J.
Chin. Med. 2011, 39, 1275–1290. [CrossRef] [PubMed]
55. Bohstam, M.; Asgary, S.; Kouhpayeh, S.; Shariati, L.; Khanhamad, H. Aptamers Against Pro- and
Anti-Inflammatory Cytokines: A Review. Inflamm. Febr. 2017, 40, 340–349. [CrossRef] [PubMed]
56. Kolehmainen, M.; Mykkänen, O.; Kirjavainen, P.V.; Leppänen, T.; Moilanen, E.; Adriaens, M.;
Laaksonen, D.E.; Hallikainen, M.; Puupponen-Pimiä, R.; Pulkkinen, L.; et al. Bilberries Reduce Low-Grade
Inflammation in Individuals with Features of Metabolic Syndrome. Mol. Nutr. Food Res. 2012, 56, 1501–1510.
[CrossRef] [PubMed]
57. Fitó, M.; Cladellas, M.; de la Torre, R.; Martí, J.; Muñoz, D.; Schröder, H.; Alcántara, M.; Pujadas-Bastardes, M.;
Marrugat, J.; Ló-Sabater, M.C.; et al. Anti-Inflammatory Effect of Virgin Olive Oil in Stable Coronary Disease
Patients: A Randomized, Crossover, Controlled Trial. Eur. J. Clin. Nutr. 2008, 62, 570–574. [CrossRef]
[PubMed]
58. Bitler, C.M.; Viale, T.M.; Damaj, B.; Crea, R. Hydrolyzed Olive Vegetation Water in Mice Has
Anti-Inflammatory Activity. J. Nutr. 2005, 135, 1475–1479. [CrossRef] [PubMed]
59. Comalada, M.; Ballester, I.; Bailon, E.; Sierra, S.; Xaus, J.; de Medina, F.; Zarzuelo, A. Inhibition of
pro-Inflammatory Markers in Primary Bone Marrow-Derived Mouse Macrophages by Naturally Occurring
Flavonoids: Analysis of the Structure-Activity Relationship. Biochem. Pharmacol. 2006, 72, 1010–1021.
[CrossRef] [PubMed]
60. Blonska, M.; Czuba, Z.P.; Krol, W. Effect of Flavone Derivatives on Interleukin-1beta (IL-1beta) MRNA
Expression and IL-1beta Protein Synthesis in Stimulated RAW 264.7 Macrophages. Scand. J. Immunol. 2003,
57, 162–166. [CrossRef] [PubMed]
61. Sharma, V.; Mishra, M.; Ghosh, S.; Tewari, R.; Basu, A.; Seth, P.; Sen, E. Modulation of Interleukin-1beta
Mediated Inflammatory Response in Human Astrocytes by Flavonoids: Implications in Neuroprotection.
Brain Res. Bull. 2007, 73, 55–63. [CrossRef] [PubMed]
62. Sato, M.; Miyazaki, T.; Kambe, F.; Maeda, K.; Seo, H. Quercetin, a Bioflavonoid, Inhibits the Induction of
Interleukin 8 and Monocyte Chemoattractant Protein-1 Expression by Tumor Necrosis Factor-Alpha in
Cultured Human Synovial Cells. J. Rheumatol. 1997, 24, 1680–1684. [PubMed]
63. Min, Y.; Choi, C.; Bark, H.; Son, H.; Park, H.; Lee, S.; Park, J.; Park, E.; Shin, H.; Kim, S. Quercetin Inhibits
Expression of Inflammatory Cytokines through Attenuation of NFkappaB and P38 MAPK in HMC-1 Human
Mast Cell Line. Inflamm. Res. 2007, 56, 210–215. [CrossRef] [PubMed]
64. Lyu, S.Y.; Park, W.B. Production of Cytokine and NO by RAW 264.7 Macrophages and PBMC in Vitro
Incubation with Flavonoids. Arch. Pharm. Res. 2005, 28, 573–581. [CrossRef] [PubMed]
65. Olivera, A.; Moore, T.W.; Hu, F.; Brown, A.P.; Sun, A.; Liotta, D.C.; Snyder, J.P.; Yoon, Y.;
Shim, H.; Marcus, A.I.; et al. Inhibition of the NF-KB Signaling Pathway by the Curcumin Analog,
3,5-Bis(2-Pyridinylmethylidene)-4-Piperidone (EF31): Anti-Inflammatory and Anti-Cancer Properties.
Int. Immunopharmacol. 2012, 12, 368–377. [CrossRef] [PubMed]
66. Drummond, E.M.; Harbourne, N.; Marete, E.; Martyn, D.; Jacquier, J.C.; O’Riordan, D.; Gibney, E.R.
Inhibition of Proinflammatory Biomarkers in THP1 Macrophages by Polyphenols Derived from Chamomile,
Meadowsweet and Willow Bark. Phyther. Res. 2013, 27, 588–594. [CrossRef] [PubMed]
67. Schindler, R.; Mancilla, J.; Endres, S.; Ghorbani, R.; Clark, S.C.; Dinarello, C.A. Correlations and Interactions in
the Production of Interleukin-6 (IL-6), IL-1, and Tumor Necrosis Factor (TNF) in Human Blood Mononuclear
Cells: IL-6 Suppresses IL-1 and TNF. Blood 1990, 75, 40–47. [PubMed]
68. Essafi-Benkhadir, K.; Refai, A.; Riahi, I.; Fattouch, S.; Karoui, H.; Essafi, M. Quince (Cydonia oblonga Miller)
Peel Polyphenols Modulate LPS-Induced Inflammation in Human THP-1-Derived Macrophages through
NF-KB, P38MAPK and Akt Inhibition. Biochem. Biophys. Res. Commun. 2012, 418, 180–185. [CrossRef]
[PubMed]
106
Nutrients 2018, 10, 1618
69. Okamoto, I.; Iwaki, K.; Koya-Miyata, S.; Tanimoto, T.; Kohno, K.; Ikeda, M.; Kurimoto, M. The Flavonoid
Kaempferol Suppresses the Graft-versus-Host Reaction by Inhibiting Type 1 Cytokine Production and
CD8+T Cell Engraftment. Clin. Immunol. 2002, 103, 132–144. [CrossRef] [PubMed]
70. Crouvezier, S.; Powell, B.; Keir, D.; Yaqoob, P. The Effects of Phenolic Components of Tea on the Production of
Pro- and Anti-Inflammatory Cytokines by Human Leukocytes in Vitro. Cytokine 2001, 13, 280–286. [CrossRef]
[PubMed]
71. Nam, N. Naturally Occurring NF-kappa B Inhibitors. Mini Rev. Med. Chem. 2006, 6, 945–951. [CrossRef]
[PubMed]
72. Hayden, M.S.; Ghosh, S. Signaling to NF-KappaB. Genes Dev. 2004, 18, 2195–2224. [CrossRef] [PubMed]
73. Haddad, J.J. Redox Regulation of pro-Inflammatory Cytokines and IkappaB-Alpha/NF-KappaB Nuclear
Translocation And. Biochem. Biophys. Res. Commun. 2002, 296, 847–856. [CrossRef]
74. Karin, M.; Ben-Neriah, Y. Phosphorylation Meets Ubiquitination: The Control of NF-[Kappa]B Activity.
Annu. Rev. Immunol. 2000, 18, 621–663. [CrossRef] [PubMed]
75. Perkins, N.D. Integrating Cell-Signalling Pathways with NF-KB and IKK Function. Nat. Rev. Mol. Cell Biol.
2007, 8, 49–62. [CrossRef] [PubMed]
76. Karin, M.; Yamamoto, Y.; Wang, Q.M. The IKK NF-KB System: A Treasure Trove for Drug Development.
Nat. Rev. Drug Discov. 2004, 3, 17–26. [CrossRef] [PubMed]
77. Rahman, I.; Biswas, S.; Kirkham, P. Regulation of Inflammation and Redox Signaling by Dietary Polyphenols.
Biochem. Pharmacol. 2006, 72, 1439–1452. [CrossRef] [PubMed]
78. Rahman, I.; Marwick, J.; Kirkham, P. Redox Modulation of Chromatin Remodeling: Impact on Histone
Acetylation and Deacetylation, NF-KappaB and pro-Inflammatory Gene Expression. Biochem. Pharmacol.
2004, 68, 1255–1267. [CrossRef] [PubMed]
79. De Stefano, D.; Maiuri, M.C.; Simeon, V.; Grassia, G.; Soscia, A.; Cinelli, M.P.; Carnuccio, R. Lycopene,
Quercetin and Tyrosol Prevent Macrophage Activation Induced by Gliadin and IFN-γ. Eur. J. Pharmacol.
2007, 566, 192–199. [CrossRef] [PubMed]
80. Comalada, M.; Camuesco, D.; Sierra, S.; Ballester, I.; Xaus, J.; Gálvez, J.; Zarzuelo, A. In Vivo
Quercitrin Anti-Inflammatory Effect Involves Release of Quercetin, Which Inhibits Inflammation through
down-Regulation of the NF-KB Pathway. Eur. J. Immunol. 2005, 35, 584–592. [CrossRef] [PubMed]
81. Ruiz, P.A.; Braune, A.; HÖlzlwimmer, G.; Quintanilla-Fend, L.; Haller, D. Quercetin Inhibits TNF-Induced
NF-KB Transcription Factor Recruitment to Proinflammatory Gene Promoters in Murine Intestinal Epithelial
Cells. J. Nutr. 2007, 137, 1208–1215. [CrossRef] [PubMed]
82. Chen, J.C.; Ho, F.M.; Chao, P.D.L.; Chen, C.P.; Jeng, K.C.G.; Hsu, H.B.; Lee, S.T.; Wen, T.W.; Lin, W.W.
Inhibition of INOS Gene Expression by Quercetin Is Mediated by the Inhibition of IκB Kinase, Nuclear
Factor-Kappa B and STAT1, and Depends on Heme Oxygenase-1 Induction in Mouse BV-2 Microglia.
Eur. J. Pharmacol. 2005, 521, 9–20. [CrossRef] [PubMed]
83. Gracia-Lafuente, A.; Guillamon, E.; Villares, A.; Rostagno, M.; Martinez, J. Flavonoids as Anti-Inflammatory
Agents: Implications in Cancer and Cardiovascular Disease. Inflamm. Res. 2009, 58, 537–552. [CrossRef]
[PubMed]
84. Nishitani, Y.; Yamamoto, K.; Yoshida, M.; Azuma, T.; Kanazawa, K.; Hashimoto, T.; Mizuno, M. Intestinal
Anti-Inflammatory Activity of Luteolin: Role of the Aglycone in NF-KB Inactivation in Macrophages
Co-Cultured with Intestinal Epithelial Cells. Biofactors 2013, 39, 522–533. [CrossRef] [PubMed]
85. Ji, G.; Zhang, Y.; Yang, Q.; Cheng, S.; Hao, J.; Zhao, X.; Jiang, Z. Genistein Suppresses LPS-Induced
Inflammatory Response through Inhibiting NF-KB Following AMP Kinase Activation in RAW 264.7
Macrophages. PLoS ONE 2012, 7, e23300870. [CrossRef] [PubMed]
86. Kim, H.H.; Bae, Y.; Kim, S.H. Galangin Attenuates Mast Cell-Mediated Allergic Inflammation. Food Chem.
Toxicol. 2013, 57, 209–216. [CrossRef] [PubMed]
87. Wheeler, D.S.; Catravas, J.D.; Odoms, K.; Denenberg, A.; Malhotra, V.; Wong, H.R. Epigallocatechin-3-Gallate,
a Green Tea-Derived Polyphenol, Inhibits IL-1 Beta-Dependent Proinflammatory Signal Transduction in
Cultured Respiratory Epithelial Cells. J. Nutr. 2004, 134, 1039–1044. [CrossRef] [PubMed]
88. Aneja, R.; Hake, P.W.; Burroughs, T.J.; Denenberg, A.G.; Wong, H.R.; Zingarelli, B. Epigallocatechin, a Green
Tea Polyphenol, Attenuates Myocardial Ischemia Reperfusion Injury in Rats. Mol. Med. 2004, 10, 55–62.
[CrossRef] [PubMed]
107
Nutrients 2018, 10, 1618
89. Ichikawa, D.; Matsui, A.; Imai, M.; Sonoda, Y.; Kasahara, T. Effect of Various Catechins on the IL-12 p40
Production by Murine Peritoneal Macrophages and A. Biol. Pharm. Bull. 2004, 27, 1353–1358. [CrossRef]
[PubMed]
90. Lin, Y.; Lin, J. Epigallocatechin-3-Gallate Blocks the Induction of Nitric Oxide Synthase by Down-Regulating
Lipopolysaccharide-Induced Activity of Transcription Factor Nuclear Factor-κB. Mol. Pharmacol. 1997, 472,
465–472. [CrossRef]
91. Mackenzie, G.; Carrasquedo, F.; Delfino, J.; Keen, C.; Fraga, C.; Oteiza, P. Epicatechin, Catechin, and Dimeric
Procyanidins Inhibit PMA? Induced NF? KappaB Activation at Multiple Steps in Jurkat T Cells. FASEB J.
2004, 18, 167–169. [CrossRef] [PubMed]
92. Carluccio, M.A.; Siculella, L.; Ancora, M.A.; Massaro, M.; Scoditti, E.; Storelli, C.; Visioli, F.; Distante, A.;
De Caterina, R. Olive Oil and Red Wine Antioxidant Polyphenols Inhibit Endothelial Activation:
Antiatherogenic Properties of Mediterranean Diet Phytochemicals. Arterioscler. Thromb. Vasc. Biol. 2003, 23,
622–629. [CrossRef] [PubMed]
93. Chang, L.; Karin, M. Mammalian MAP Kinase Signalling Cascades. Nature 2001, 410, 37–40. [CrossRef]
[PubMed]
94. Khan, N.; Afaq, F.; Saleem, M.; Ahmad, N.; Mukhtar, H. Targeting Multiple Signaling Pathways by Green
Tea Polyphenol (−)-Epigallocatechin-3-Gallate. 1 Khan N, Afaq F, Saleem M, Ahmad N, Mukhtar H. Author
Information Full. Cancer Res. 2006, 66, 2500–2505. [CrossRef] [PubMed]
95. Kolch, W. Coordinating ERK/MAPK Signalling through Scaffolds and Inhibitors. Nat. Rev. Mol. Cell Biol.
2005, 6, 827–837. [CrossRef] [PubMed]
96. Lu, Z.; Xu, S. ERK1/2 MAP Kinases in Cell Survival and Apoptosis. IUBMB Life 2006, 58, 621–631. [CrossRef]
[PubMed]
97. Mayor, F.; Jurado-Pueyo, M.; Campos, P.M.; Murga, C. Interfering with MAP Kinase Docking Interactions:
Implications and Perspective for the P38 Route. Cell Cycle 2007, 6, 528–533. [CrossRef] [PubMed]
98. Kaminska, B. MAPK Signalling Pathways as Molecular Targets for Anti-Inflammatory Therapy—From
Molecular Mechanisms to Therapeutic Benefits. Biochim. Biophys. Acta 2005, 1754, 253–262. [CrossRef]
[PubMed]
99. Karin, M. Inflammation-Activated Protein Kinases as Targets for Drug Development. Proc. Am. Thorac. Soc.
2005, 2, 386–390. [CrossRef] [PubMed]
100. Xagorari, A.; Roussos, C.; Papapetropoulos, A. Inhibition of LPS-Stimulated Pathways in Macrophages by
the Flavonoid Luteolin. Br. J. Pharmacol. 2002, 136, 1058–1064. [CrossRef] [PubMed]
101. Chen, C.; Chow, M.; Huang, W.; Lin, Y.; Chang, Y. Flavonoids Inhibit Tumor Necrosis Factor-Alpha-Induced
up-Regulation of Intercellular Adhesion Molecule-1 (ICAM-1) in Respiratory Epithelial Cells through
Activator Protein-1 and Nuclear Factor-KappaB: Structure-Activity Relationships. Mol. Pharmacol. 2004, 66,
683–693. [PubMed]
102. Wadsworth, T.L.; McDonald, T.L.; Koop, D.R. Effects of Ginkgo Biloba Extract (EGb 761) and Quercetin on
Lipopolysaccharide-Induced Signaling Pathways Involved in the Release of Tumor Necrosis Factor-Alpha.
Biochem. Pharmacol. 2001, 62, 963–974. [CrossRef]
103. Cho, S.; Park, S.; Kwon, M.; Jeong, T.; Bok, S.; Choi, W.; Jeong, W.; Ryu, S.; Do, S.; Song, C.; et al. Quercetin
Suppresses Proinflammatory Cytokines Production through MAP Kinases AndNF-Kappa B Pathway in
Lipopolysaccharide-Stimulated Macrophage. Mol. Cell. Biochem. 2003, 243, 153–160. [CrossRef] [PubMed]
104. Kundu, J.K.; Surh, Y.J. Epigallocatechin Gallate Inhibits Phorbol Ester-Induced Activation of NF-KB and
CREB in Mouse Skin Role of P38 MAPK. Ann. N. Y. Acad. Sci. 2007, 1095, 504–512. [CrossRef] [PubMed]
105. Pasten, C.; Olave, N.; Zhou, L.; Tabengwa, E.; Wolkowicz, P.; Grenett, H. Polyphenols Downregulate
PAI-1 Gene Expression in Cultured Human Coronary Artery Endothelial Cells: Molecular Contributor to
Cardiovascular Protection. Thromb. Res. 2007, 121, 59–65. [CrossRef] [PubMed]
106. Chandrasekharan, N.V.; Dai, H.; Roos, K.L.T.; Evanson, N.K.; Tomsik, J.; Elton, T.S.; Simmons, D.L. COX-3,
a Cyclooxygenase-1 Variant Inhibited by Acetaminophen and Other Analgesic/Antipyretic Drugs: Cloning,
Structure, and Expression. Proc. Natl. Acad. Sci. USA 2002, 99, 13926–13931. [CrossRef] [PubMed]
107. Needleman, P.; Isakson, P. The Discovery and Function of COX-2. J. Rheumatol. Suppl. 2018, 49, 6–8.
108. Kim, H.P.; Son, K.H.; Chang, H.W.; Kang, S.S. Anti-Inflammatory Plant Flavonoids and Cellular Action
Mechanisms. J. Pharmacol. Sci. 2004, 96, 229–245. [CrossRef] [PubMed]
108
Nutrients 2018, 10, 1618
109. Welton, A.F.; Tobias, L.D.; Fiedler-Nagy, C.; Anderson, W.; Hope, W.; Meyers, K.; Coffey, J.W. Effect of
Flavonoids on Arachidonic Acid Metabolism. Prog. Clin. Biol. Res. 1986, 213, 231–242. [PubMed]
110. Laughton, M.; Evans, P.; Moroney, M.; Hoult, J.; Halliwell, B. Inhibition of Mammalian 5-Lipoxygenase and
Cyclo-Oxygenase by Flavonoids and Phenolic Dietary Additives. Relationship to Antioxidant Activity and
to Iron Ion-Reducing Ability. Biochem. Pharmacol. 1991, 42, 1673–1681. [CrossRef]
111. Aviram, M.; Fuhrman, B. Polyphenolic Flavonoids Inhibit Macrophage-Mediated Oxidation of LDL and
Attenuate Atherogenesis. Atherosclerosis 1998, 137, 9694541. [CrossRef]
112. Ferrandiz, M.L.; Alcaraz, M.J. Ferrandiz 1991-Anti-Inflammatory Activity and Inhibition of Arachidonic
Acid Metabolism by Flavonoids. Agent Action 1991, 32, 283–288. [CrossRef]
113. Kim, H.; Mani, I.; Iversen, L.; Ziboh, V. Effects of Naturally-Occurring Flavonoids and Biflavonoids on
Epidermal Cyclooxygenase and Lipoxygenase from Guinea-Pigs. Prostaglandin Leukot. Essent. Fat. Acid.
1998, 58, 17–24. [CrossRef]
114. Luceri, C.; Caderni, G.; Sanna, A.; Dolara, P. Red Wine and Black Tea Polyphenols Modulate the
Expression of Cycloxygenase-2, Inducible Nitric Oxide Synthase and Glutathione-Related Enzymes in
Azoxymethane-Induced F344 Rat Colon Tumors. J. Nutr. 2002, 132, 1376–1379. [CrossRef] [PubMed]
115. Hou, D.X.; Luo, D.; Tanigawa, S.; Hashimoto, F.; Uto, T.; Masuzaki, S.; Fujii, M.; Sakata, Y. Prodelphinidin
B-4 3 -O-Gallate, a Tea Polyphenol, Is Involved in the Inhibition of COX-2 and INOS via the Downregulation
of TAK1-NF-KB Pathway. Biochem. Pharmacol. 2007, 74, 742–751. [CrossRef] [PubMed]
116. Hou, D.; Masuzaki, S.; Hashimoto, F.; Uto, T.; Tanigawa, S.; Fujii, M.; Sakata, Y. Green Tea Proanthocyanidins
Inhibit Cyclooxygenase-2 Expression in LPS-Activated Mouse Macrophages: Molecular Mechanisms and
Structure? Activity Relationship. Arch. Biochem. Biophys. 2007, 460, 67–74. [CrossRef] [PubMed]
117. Miles, E.A.; Zoubouli, P.; Calder, P.C. Differential Anti-Inflammatory Effects of Phenolic Compounds from
Extra Virgin Olive Oil Identified in Human Whole Blood Cultures. Nutrition 2005, 21, 389–394. [CrossRef]
[PubMed]
118. Tuck, K.L.; Hayball, P.J. Major Phenolic Compounds in Olive Oil: Metabolism and Health Effects.
J. Nutr. Biochem. 2002, 13, 636–644. [CrossRef]
119. De la Puerta, R.; Gutierrez, V.R.; Hoult, J. Inhibition of Leukocyte 5 Lipoxygenase by Phenolics from Virgin
Olive Oil. Biochem. Pharmacol. 1999, 57, 445–449. [CrossRef]
120. Beauchamp, G.K.; Keast, R.S.J.; Morel, D.; Lin, J.; Pika, J.; Han, Q.; Lee, C.H.; Smith, A.B.; Breslin, P.A.S.
Ibuprofen-like Activity in Extra-Virgin Olive Oil. Nature 2005, 437, 45–46. [CrossRef] [PubMed]
121. Berlett, B.S.; Stadtman, E.R.; Berlett, B.S.; Stadtman, E.R. Protein Oxidation in Aging, Disease, and Oxidative
Stress. J. Biol. Chem. 1997, 272, 20313–20316. [CrossRef] [PubMed]
122. Salzano, S.; Checconi, P.; Hanschmann, E.-M.; Lillig, C.H.; Bowler, L.D.; Chan, P.; Vaudry, D.; Mengozzi, M.;
Coppo, L.; Sacre, S.; et al. Linkage of Inflammation and Oxidative Stress via Release of Glutathionylated
Peroxiredoxin-2, Which Acts as a Danger Signal. Proc. Natl. Acad. Sci. USA 2014, 111, 12157–12162.
[CrossRef] [PubMed]
123. Willcox, J.K.; Ash, S.L.; Catignani, G.L. Antioxidants and Prevention of Chronic Disease. Crit. Rev. Food Sci.
Nutr. 2004, 44, 275–295. [CrossRef] [PubMed]
124. Bryan, N.; Ahswin, H.; Smart, N.; Bayon, Y.; Wohlert, S.; Hunt, J.A. Reactive Oxygen Species (ROS)-A Family
of Fate Deciding Molecules Pivotal in Constructive Inflammation and Wound Healing. Eur. Cells Mater. 2012,
24, 249–265. [CrossRef]
125. Naik, E.; Dixit, V.M. Mitochondrial Reactive Oxygen Species Drive Proinflammatory Cytokine Production:
Figure 1. J. Exp. Med. 2011, 208, 417–420. [CrossRef] [PubMed]
126. Clark, R.A.; Valente, A.J. Nuclear Factor Kappa B Activation by NADPH Oxidases. Mech. Ageing Dev. 2004,
125, 799–810. [CrossRef] [PubMed]
127. Geiszt, M.; Leto, T.L. The Nox Family of NAD(P)H Oxidases: Host Defense and Beyond. J. Biol. Chem. 2004,
279, 51715–51718. [CrossRef] [PubMed]
128. Heim, K.E.; Tagliaferro, A.R.; Bobilya, D.J. Flavonoid Antioxidants: Chemistry, Metabolism and
Structure-Activity Relationships. J. Nutr. Biochem. 2002, 13, 572–584. [CrossRef]
129. Mishra, A.; Sharma, A.K.; Kumar, S.; Saxena, A.K.; Pandey, A.K. Bauhinia Variegata Leaf Extracts Exhibit
Considerable Antibacterial, Antioxidant, and Anticancer Activities. Biomed. Res. Int. 2013, 2013, 915436.
[CrossRef] [PubMed]
109
Nutrients 2018, 10, 1618
130. Mishra, A.; Kumar, S.; Pandey, A.K. Scientific Validation of the Medicinal Efficacy of Tinospora Cordifolia.
Sci. World J. 2013, 2013, 292934. [CrossRef] [PubMed]
131. Marnett, L.J.; Riggins, J.N.; West, J.D. Endogenous Generation of Reactive Oxidants and Electrophiles and
Their Reactions with DNA and Protein. J. Clin. Investig. 2003, 111, 583–593. [CrossRef] [PubMed]
132. Prousek, J. Fenton Chemistry in Biology and Medicine. Pure Appl. Chem. 2007, 79, 2007–2010. [CrossRef]
133. Deby-Dupont, G.; Mouithys-Mickalad, A.; Serteyn, D.; Lamy, M.; Deby, C. Resveratrol and Curcumin Reduce
the Respiratory Burst of Chlamydia-Primed THP-1 Cells. Biochem. Biophys. Res. Commun. 2005, 333, 21–27.
[CrossRef] [PubMed]
134. Chow, S.E.; Hshu, Y.C.; Wang, J.S.; Chen, J.K. Resveratrol Attenuates OxLDL-Stimulated NADPH Oxidase
Activity and Protects Endothelial Cells from Oxidative Functional Damages. J. Appl. Physiol. 2007, 102,
1520–1527. [CrossRef] [PubMed]
135. Petrônio, M.S.; Zeraik, M.L.; Da Fonseca, L.M.; Ximenes, V.F. Apocynin: Chemical and Biophysical Properties
of a NADPH Oxidase Inhibitor. Molecules 2013, 18, 2821–2839. [CrossRef] [PubMed]
136. Shen, L.; Ji, H.F. Insights into the Inhibition of Xanthine Oxidase by Curcumin. Bioorg. Med. Chem. Lett. 2009,
19, 5990–5993. [CrossRef] [PubMed]
137. Aucamp, J. Inhibition of Xanthine Oxidase by Tea Catechins (Camellia Sinensis). Method Mol. Biol. 1997, 702,
47–60.
138. Schmidt, A.; Böhmer, A.E.; Antunes, C.; Schallenberger, C.; Porciuncula, L.; Elisabetsky, E.; Lara, D.; Souza, D.
Anti-Nociceptive Properties of the Xanthine Oxidase Inhibitor Allopurinol in Mice: Role of A1 Adenosine
Receptors. Br. J. Pharmacol. 2009, 156, 163–172. [CrossRef] [PubMed]
139. Nguyen, M.T.T.; Nguyen, N.T. Xanthine Oxidase Inhibitors from Vietnamese Blume balsamifer L. Phyther. Res.
2012, 26, 1178–1181. [CrossRef] [PubMed]
140. Bräunlich, M.; Slimestad, R.; Wangensteen, H.; Brede, C.; Malterud, K.E.; Barsett, H. Extracts, Anthocyanins
and Procyanidins from Aronia Melanocarpa as Radical Scavengers and Enzyme Inhibitors. Nutrients 2013, 5,
663–678. [CrossRef] [PubMed]
141. Huang, X.F.; Li, H.Q.; Shi, L.; Xue, J.Y.; Ruan, B.F.; Zhu, H.L. Synthesis of Resveratrol Analogues, and
Evaluation of Their Cytotoxic and Xanthine Oxidase Inhibitory Activities. Chem. Biodivers. 2008, 5, 636–642.
[CrossRef] [PubMed]
142. Cheon, B.S.; Kim, Y.H.; Son, K.S.; Chang, H.W.; Kang, S.S.; Kim, H.P. Effects of Prenylated Flavonoids and
Biflavonoids on Lipopolysaccharide-Induced Nitric Oxide Production from the Mouse Macrophage Cell
Line RAW 264.7. Planta Med. 2000, 66, 596–600. [CrossRef] [PubMed]
143. Sarkar, A.; Bhaduri, A. Black Tea Is a Powerful Chemopreventor of Reactive Oxygen and Nitrogen Species:
Comparison with Its Individual Catechin Constituents and Green Tea. Biochem. Biophys. Res. Commun. 2001,
284, 173–178. [CrossRef] [PubMed]
144. Sporn, M.B.; Liby, K.T. NRF2 and Cancer: The Good, the Bad and the Importance of Context. Nat. Rev. Cancer
2012, 12, 564–571. [CrossRef] [PubMed]
145. Chu, A. Antagonism by Bioactive Polyphenols Against Inflammation: A Systematic View. Inflamm. Allergy
Drug Targets 2014, 13, 34–64. [CrossRef] [PubMed]
146. Meydani, M.; Hasan, S.T. Dietary Polyphenols and Obesity. Nutrients 2010, 2, 737–751. [CrossRef] [PubMed]
147. Yahfoufi, N.; Mallet, J.F.; Graham, E.; Matar, C. Role of Probiotics and Prebiotics in Immunomodulation.
Curr. Opin. Food Sci. 2018, 20, 82–91. [CrossRef]
148. Roy, D.; Perreault, M.; Marette, A. Insulin Stimulation of Glucose Uptake in Skeletal Muscles and Adipose
Tissues in Vivo Is NO Dependent. Am. J. Physiol. Endocrinol. Metab. 1998, 274, E692–E699. [CrossRef]
149. Fryer, L.G.; Hajduch, E.; Rencurel, F.; Salt, I.P.; Hundal, H.S.; Hardie, D.G.; Carling, D. Activation of Glucose
Transport by AMP-Activated Protein Kinase via Stimulation of Nitric Oxide Synthase. Diabetes 2000, 49,
1978–1985. [CrossRef] [PubMed]
150. Roberts, C.K.; Barnard, R.J.; Scheck, S.H.; Balon, T.W. Exercise-Stimulated Glucose Transport in Skeletal
Muscle Is Nitric Oxide Dependent. Am. J. Physiol. 1997, 273, E220–E225. [PubMed]
151. Peters, U.; Poole, C.; Arab, L. Does Tea Affect Cardiovascular Disease? A Meta-Analysis. Am. J. Epidemiol.
2001, 154, 495–503. [CrossRef] [PubMed]
152. Lindsay, J.; Laurin, D.; Verreault, R.; Hébert, R.; Helliwell, B.; Hill, G.; McDowell, I. Risk Factors for Alzheimer’s
Disease: A Prospective Analysis from the Canadian Study of Health and Aging. Am. J. Epidemiol. 2002, 156,
445–453. [CrossRef] [PubMed]
110
Nutrients 2018, 10, 1618
153. Truelsen, T.; Thudium, D.; Grønbaek, M.; Copenhagen City Heart Study. Amount and Type of Alcohol and
Risk of Dementia: The Copenhagen City Heart Study. Neurology 2002, 59, 1313–1319. [CrossRef] [PubMed]
154. Hadi, S.M.; Asad, S.F.; Singh, S.; Ahmad, A. Putative Mechanism for Anticancer and Apoptosis-Inducing
Properties of Plant-Derived Polyphenolic Compounds. IUBMB Life 2000, 50, 167–171. [PubMed]
155. Park, S.; Ahmad, F.; Philip, A.; Baar, K.; William, T.; Luo, H.; Ke, H.; Rehmann, H.; Taussing, R.; Brown, A.;
et al. Resveratrol Ameliorates Aging-Related Metabolic Phenotypes by Inhibiting CAMP Phosphodiesterases.
Cell 2012, 148, 421–433. [CrossRef] [PubMed]
156. Wallerath, T.; Deckert, G.; Ternes, T.; Anderson, H.; Li, H.; Witte, K.; Forstermann, U. Resveratrol,
a Polyphenolic Phytoalexin Present in Red Wine, Enhances Expression and Activity of Endothelial Nitric
Oxide Synthase. Circulation 2002, 106, 1652–1658. [CrossRef] [PubMed]
157. Kumar, S.; Narwal, S.; Kumar, V.; Prakash, O. α-Glucosidase Inhibitors from Plants: A Natural Approach to
Treat Diabetes. Pharmacogn. Rev. 2011, 5, 19–29. [CrossRef] [PubMed]
158. Di Castelnuovo, A.; Rotondo, S.; Iacoviello, L.; Donati, M.B.; De Gaetano, G. Meta-Analysis of Wine and Beer
Consumption in Relation to Vascular Risk. Circulation 2002, 105, 2836–2844. [CrossRef] [PubMed]
159. Hooper, L.; Kroon, P.A.; Rimm, E.B.; Cohn, J.S.; Harvey, I.; Cornu, K.A.; Le Ryder, J.J.; Hall, W.L.; Cassidy, A.
Flavonoids, Flavonoid-Rich Foods, and Cardiovascular Risk: A Meta-Analysis of Randomized Controlled
Trials 1, 2. Am. J. Clin. Nutr. 2008, 88, 38–50. [CrossRef] [PubMed]
160. Shen, M.; Zhao, L.; Wu, R.X.; Yue, S.Q.; Pei, J.M. The Vasorelaxing Effect of Resveratrol on Abdominal Aorta
from Rats and Its Underlying Mechanisms. Vasc. Pharmacol. 2013, 58, 64–70. [CrossRef] [PubMed]
161. Peppa, M.; Raptis, S.A. Advanced Glycation End Products and Cardiovascular Disease. Curr. Diabete Rev.
2008, 4, 92–100. [CrossRef]
162. Huang, S.M.; Wu, C.H.; Yen, G.C. Effects of Flavonoids on the Expression of the Pro-Inflammatory Response
in Human Monocytes Induced by Ligation of the Receptor for AGEs. Mol. Nutr. Food Res. 2006, 50, 1129–1139.
[CrossRef] [PubMed]
163. Kim, J.M.; Lee, E.K.; Kim, D.H.; Yu, B.P.; Chung, H.Y. Kaempferol Modulates Pro-Inflammatory NF-KB
Activation by Suppressing Advanced Glycation Endproducts-Induced NADPH Oxidase. Age 2010, 32,
197–208. [CrossRef] [PubMed]
164. Wilkinson-Berka, J.L.; Rana, I.; Armani, R.; Agrotis, A. Reactive Oxygen Species, Nox and Angiotensin II in
Angiogenesis: Implications for Retinopathy. Clin. Sci. 2013, 124, 597–615. [CrossRef] [PubMed]
165. Thomasset, S.; Teller, N.; Cai, H.; Marko, D.; Berry, D.; Steward, W.; Gescher, A. Do Anthocyanins and
Anthocyanidins, Cancer Chemopreventive Pigments in the Diet, Merit Development as Potential Drugs?
Cancer Chemother. Pharmacol. 2009, 64, 201–211. [CrossRef] [PubMed]
166. Aviram, M.; Fuhrman, B. Wine Flavonoids Protect against LDL Oxidation and Atherosclerosis. Ann. N. Y.
Acad. Sci. 2002, 957, 146–161. [CrossRef] [PubMed]
167. Commenges, D.; Scotet, V.; Renaud, S.; Jacqmin-Gadda, H.; Barberger-Gateau, P.; Dartigues, J.F. Intake of
Flavonoids and Risk of Dementia. Eur. J. Epidemiol. 2000, 16, 357–363. [CrossRef] [PubMed]
168. Dai, Q.; Borenstein, A.R.; Wu, Y.; Jackson, J.C.; Larson, E.B. Fruit and Vegetable Juices and Alzheimer’s
Disease: The Kame Project. Am. J. Med. 2006, 119, 751–759. [CrossRef] [PubMed]
169. Morris, M.C.; Evans, D.A.; Tangney, C.C.; Bienias, J.L.; Wilson, R.S. Associations of Vegetable and Fruit
Consumption with Age-Related Cognitive Change. Neurology 2006, 67, 1370–1376. [CrossRef] [PubMed]
170. Checkoway, H.; Powers, K.; Smith-Weller, T.; Franklin, G.M.; Longstreth, W.T.; Swanson, P.D. Parkinson’s
Disease Risks Associated with Cigarette Smoking, Alcohol Consumption, and Caffeine Intake. Am. J. Epidemiol.
2002, 155, 732–738. [CrossRef] [PubMed]
171. Shehzad, A.; Lee, Y.S. Molecular Mechanisms of Curcumin Action: Signal Transduction. Biofactors 2013, 39,
27–36. [CrossRef] [PubMed]
172. Gomez-Pinilla, F.; Nguyen, T.T.J. Natural Mood Foods: The Actions of Polyphenols against Psychiatric and
Cognitive Disorders. Nutr. Neurosci. 2012, 15, 127–133. [CrossRef] [PubMed]
173. Vauzour, D.; Vafeiadou, K.; Rice-Evans, C.; Williams, R.J.; Spencer, J.P.E. Activation of Pro-Survival Akt
and ERK1/2 Signalling Pathways Underlie the Anti-Apoptotic Effects of Flavanones in Cortical Neurons.
J. Neurochem. 2007, 103, 1355–1367. [CrossRef] [PubMed]
174. Vafeiadou, K.; Vauzour, D.; Lee, H.Y.; Rodriguez-Mateos, A.; Williams, R.J.; Spencer, J.P.E. The Citrus Flavanone
Naringenin Inhibits Inflammatory Signalling in Glial Cells and Protects against Neuroinflammatory Injury.
Arch. Biochem. Biophys. 2009, 484, 100–109. [CrossRef] [PubMed]
111
Nutrients 2018, 10, 1618
175. Wang, X.; Chen, S.; Ma, G.; Ye, M.; Lu, G. Genistein Protects Dopaminergic Neurons by Inhibiting Microglial
Activation. Neuroreport 2005, 16, 267–270. [CrossRef] [PubMed]
176. Bhat, N.R.; Feinstein, D.L.; Shen, Q.; Bhat, A.N. P38 MAPK-Mediated Transcriptional Activation of Inducible
Nitric-Oxide Synthase in Glial Cells: Roles of Nuclear Factors, Nuclear Factor KB, CAMP Response
Element-Binding Protein, CCAAT/Enhancer-Binding Protein-β, and Activating Transcription Factor-2.
J. Biol. Chem. 2002, 277, 29584–29592. [CrossRef] [PubMed]
177. Whiting, S.; Derbyshire, E.; Tiwari, B.K. Capsaicinoids and Capsinoids. A Potential Role for Weight
Management? A Systematic Review of the Evidence. Appetite 2012, 59, 341–348. [CrossRef] [PubMed]
178. Saito, M.; Yoneshiro, T. Capsinoids and Related Food Ingredients Activating Brown Fat Thermogenesis and
Reducing Body Fat in Humans. Curr. Opin. Lipidol. 2013, 24, 71–77. [CrossRef] [PubMed]
179. Higuchi, M.; Dusting, G.J.; Peshavariya, H.; Jiang, F.; Hsiao, S.T.-F.; Chan, E.C.; Liu, G.-S. Differentiation
of Human Adipose-Derived Stem Cells into Fat Involves Reactive Oxygen Species and Forkhead Box O1
Mediated Upregulation of Antioxidant Enzymes. Stem Cell Dev. 2013, 22, 878–888. [CrossRef] [PubMed]
180. Okamoto, M.; Irii, H.; Tahara, Y.; Ishii, H.; Hirao, A.; Udagawa, H.; Hiramoto, M.; Yasuda, K.; Takanishi, A.;
Shibata, S.; et al. Synthesis of a New [6]-Gingerol Analogue and Its Protective Effect with Respect to the
Development of Metabolic Syndrome in Mice Fed a High-Fat Diet. J. Med. Chem. 2011, 54, 6295–6304.
[CrossRef] [PubMed]
181. Panahi, Y.; Hosseini, M.S.; Khalili, N.; Naimi, E.; Soflaei, S.S.; Majeed, M.; Sahebkar, A. Effects of
Supplementation with Curcumin on Serum Adipokine Concentrations: A Randomized Controlled Trial.
Nutrition 2016, 32, 1116–1122. [CrossRef] [PubMed]
182. Yang, C.S.; Landau, J.M.; Huang, M.T.; Newmark, H.L. Inhibition of Carcinogenesis by Dietary Polyphenolic
Compounds. Annu. Rev. Nutr. 2001, 21, 381–406. [CrossRef] [PubMed]
183. Wenzel, U.; Kuntz, S.; Brendel, M.D.; Daniel, H. Dietary Flavone Is a Potent Apoptosis Inducer in Human
Colon Carcinoma Cells. Cancer Res. 2000, 60, 3823–3831. [PubMed]
184. Turrini, E.; Ferruzzi, L.; Fimognari, C. Potential Effects of Pomegranate Polyphenols in Cancer Prevention
and Therapy. Oxid. Med. Cell. Longev. 2015, 2015, 938475. [CrossRef] [PubMed]
185. Wessner, B.; Strasser, E.-M.; Koitz, N.; Schmuckenschlager, C.; Unger-Manhart, N.; Roth, E. Green Tea
Polyphenol Administration Partly Ameliorates Chemotherapy-Induced Side Effects in the Small Intestine of
Mice. J. Nutr. 2007, 137, 634–640. [CrossRef] [PubMed]
186. Harper, C.E.; Patel, B.B.; Wang, J.; Eltoum, I.A.; Lamartiniere, C.A. Epigallocatechin-3-Gallate Suppresses
Early Stage, but Not Late Stage Prostate Cancer in TRAMP Mice: Mechanisms of Action. Prostate 2007, 67,
1576–1589. [CrossRef] [PubMed]
187. Chuang, S.E.; Cheng, A.L.; Lin, J.K.; Kuo, M.L. Inhibition by Curcumin of Diethylnitrosamine-Induced
Hepatic Hyperplasia, Inflammation, Cellular Gene Products and Cell-Cycle-Related Proteins in Rats.
Food Chem. Toxicol. 2000, 38, 991–995. [CrossRef]
188. Link, A.; Balaguer, F.; Goel, A. Cancer Chemoprevention by Dietary Polyphenols: Promising Role for
Epigenetics. Biochem. Pharmacol. 2010, 80, 1771–1792. [CrossRef] [PubMed]
189. Brenner, D.E.; Gescher, A.J. Cancer Chemoprevention: Lessons Learned and Future Directions. Br. J. Cancer
2005, 93, 735–739. [CrossRef] [PubMed]
190. Weng, C.-J.; Yen, G.-C. Chemopreventive Effects of Dietary Phytochemicals against Cancer Invasion and
Metastasis: Phenolic Acids, Monophenol, Polyphenol, and Their Derivatives. Cancer Treat. Rev. 2012, 38,
76–87. [CrossRef] [PubMed]
191. Liou, G.-Y.; Storz, P. Reactive Oxygen Species in Cancer. Free Radic. Res. 2010, 44, 479–496. [CrossRef]
[PubMed]
192. Wang, C.; Schuller Levis, G.B.; Lee, E.B.; Levis, W.R.; Lee, D.W.; Kim, B.S.; Park, S.Y.; Park, E. Platycodin D
and D3 Isolated from the Root of Platycodon Grandiflorum Modulate the Production of Nitric Oxide and
Secretion of TNF-Alpha in Activated RAW 264.7 Cells. Int. Immunopharmacol. 2004, 4, 1039–1049. [CrossRef]
[PubMed]
193. Amararathna, M.; Johnston, M.R.; Rupasinghe, H.P.V. Plant Polyphenols as Chemopreventive Agents for
Lung Cancer. Int. J. Mol. Sci. 2016, 17, 1352. [CrossRef] [PubMed]
194. Tsuji, P.A.; Walle, T. Inhibition of Benzo[a]Pyrene-Activating Enzymes and DNA Binding in Human Bronchial
Epithelial BEAS-2B Cells by Methoxylated Flavonoids. Carcinogenesis 2006, 27, 1579–1585. [CrossRef]
[PubMed]
112
Nutrients 2018, 10, 1618
195. Zhai, X.; Lin, M.; Zhang, F.; Hu, Y.; Xu, X.; Li, Y.; Liu, K.; Ma, X.; Tian, X.; Yao, J. Dietary Flavonoid Genistein
Induces Nrf2 and Phase II Detoxification Gene Expression via ERKs and PKC Pathways and Protects against
Oxidative Stress in Caco-2 Cells. Mol. Nutr. Food Res. 2013, 57, 249–259. [CrossRef] [PubMed]
196. Lambert, J.D.; Elias, R.J. The Antioxidant and Pro-Oxidant Activities of Green Tea Polyphenols: A Role in
Cancer Prevention. Arch. Biochem. Biophys. 2010, 501, 65–72. [CrossRef] [PubMed]
197. Nakazato, T.; Ito, K.; Ikeda, Y.; Kizaki, M. Green Tea Component, Catechin, Induces Apoptosis of Human
Malignant B Cells via Production of Reactive Oxygen Species. Clin. Cancer Res. 2005, 11, 6040–6049.
[CrossRef] [PubMed]
198. Howells, L.M.; Mitra, A.; Manson, M.M. Comparison of Oxaliplatin- and Curcumin-Mediated
Antiproliferative Effects in Colorectal Cell Lines. Int. J. Cancer 2007, 121, 175–183. [CrossRef] [PubMed]
199. Balasubramanian, S.; Efimova, T.; Eckert, R.L. Green Tea Polyphenol Stimulates a Ras, MEKK1, MEK3,
and P38 Cascade to Increase Activator Protein 1 Factor-Dependent Involucrin Gene Expression in Normal
Human Keratinocytes. J. Biol. Chem. 2002, 277, 1828–1836. [CrossRef] [PubMed]
200. Kao, Y.-L.; Kuo, Y.-M.; Lee, Y.-R.; Yang, S.-F.; Chen, W.-R.; Lee, H.-J. Apple Polyphenol Induces Cell Apoptosis,
Cell Cycle Arrest at G2/M Phase, and Mitotic Catastrophe in Human Bladder Transitional Carcinoma Cells.
J. Funct. Food 2015, 14, 384–394. [CrossRef]
201. Singh, M.; Singh, R.; Bhui, K.; Tyagi, S.; Mahmood, Z.; Shukla, Y. Tea Polyphenols Induce Apoptosis through
Mitochondrial Pathway and by Inhibiting Nuclear Factor-KappaB and Akt Activation in Human Cervical
Cancer Cells. Oncol. Res. 2011, 19, 245–257. [CrossRef] [PubMed]
202. Monasterio, A.; Urdaci, M.C.; Pinchuk, I.V.; López-Moratalla, N.; Martínez-Irujo, J.J. Flavonoids Induce
Apoptosis in Human Leukemia U937 Cells through Caspase- and Caspase-Calpain-Dependent Pathways.
Nutr. Cancer 2004, 50, 90–100. [CrossRef] [PubMed]
203. Brusselmans, K.; Vrolix, R.; Verhoeven, G.; Swinnen, J.V. Induction of Cancer Cell Apoptosis by Flavonoids
Is Associated with Their Ability to Inhibit Fatty Acid Synthase Activity. J. Biol. Chem. 2005, 280, 5636–5645.
[CrossRef] [PubMed]
204. Lee, S.H.; Yumnam, S.; Hong, G.E.; Raha, S.; Saralamma, V.V.G.; Lee, H.J.; Heo, J.D.; Lee, S.J.; Lee, W.-S.;
Kim, E.-H.; et al. Flavonoids of Korean Citrus Aurantium L. Induce Apoptosis via Intrinsic Pathway in
Human Hepatoblastoma HepG2 Cells. Phyther. Res. PTR 2015, 29, 1940–1949. [CrossRef] [PubMed]
205. Castillo-Pichardo, L.; Dharmawardhane, S.F. Grape Polyphenols Inhibit Akt/Mammalian Target of
Rapamycin Signaling and Potentiate the Effects of Gefitinib in Breast Cancer. Nutr. Cancer 2012, 64, 1058–1069.
[CrossRef] [PubMed]
206. Sepporta, M.V.; Fuccelli, R.; Rosignoli, P.; Ricci, G.; Servili, M.; Morozzi, G.; Fabiani, R. Oleuropein Inhibits
Tumour Growth and Metastases Dissemination in Ovariectomised Nude Mice with MCF-7 Human Breast
Tumour Xenografts. J. Funct. Food 2014, 8, 269–273. [CrossRef]
207. Rivera, A.R.; Castillo-Pichardo, L.; Gerena, Y.; Dharmawardhane, S. Anti-Breast Cancer Potential of Quercetin
via the Akt/AMPK/Mammalian Target of Rapamycin (MTOR) Signaling Cascade. PLoS ONE 2016, 11,
e0157251. [CrossRef] [PubMed]
208. Xia, Y.; Shen, S.; Verma, I.M. NF-KB, an Active Player in Human Cancers. Cancer Immunol. Res. 2014, 2,
823–830. [CrossRef] [PubMed]
209. Kim, J.-M.; Noh, E.-M.; Kwon, K.-B.; Kim, J.-S.; You, Y.-O.; Hwang, J.-K.; Hwang, B.-M.; Kim, B.-S.; Lee, S.-H.;
Lee, S.J.; et al. Curcumin Suppresses the TPA-Induced Invasion through Inhibition of PKCα-Dependent
MMP-Expression in MCF-7 Human Breast Cancer Cells. Phytomed. Int. J. Phyther. Phytopharm. 2012, 19,
1085–1092. [CrossRef] [PubMed]
210. Sarkar, F.H.; Li, Y.; Wang, Z.; Kong, D. The Role of Nutraceuticals in the Regulation of Wnt and Hedgehog
Signaling in Cancer. Cancer Metastasis Rev. 2010, 29, 383–394. [CrossRef] [PubMed]
211. Aggarwal, B.B. Nuclear Factor-KappaB: The Enemy Within. Cancer Cell 2004, 6, 203–208. [CrossRef]
[PubMed]
212. Bachmeier, B.; Nerlich, A.G.; Iancu, C.M.; Cilli, M.; Schleicher, E.; Vené, R.; Dell’Eva, R.; Jochum, M.;
Albini, A.; Pfeffer, U. The Chemopreventive Polyphenol Curcumin Prevents Hematogenous Breast Cancer
Metastases in Immunodeficient Mice. Cell. Physiol. Biochem. 2007, 19, 137–152. [CrossRef] [PubMed]
213. Farhangi, B.; Alizadeh, A.M.; Khodayari, H.; Khodayari, S.; Dehghan, M.J.; Khori, V.; Heidarzadeh, A.;
Khaniki, M.; Sadeghiezadeh, M.; Najafi, F. Protective Effects of Dendrosomal Curcumin on an Animal
Metastatic Breast Tumor. Eur. J. Pharmacol. 2015, 758, 188–196. [CrossRef] [PubMed]
113
Nutrients 2018, 10, 1618
214. Tsai, J.H.; Yang, J. Epithelial–Mesenchymal Plasticity in Carcinoma Metastasis. Gene Dev. 2013, 27, 2192–2206.
[CrossRef] [PubMed]
215. Kang, J.; Kim, E.; Kim, W.; Seong, K.M.; Youn, H.; Kim, J.W.; Kim, J.; Youn, B. Rhamnetin and Cirsiliol
Induce Radiosensitization and Inhibition of Epithelial-Mesenchymal Transition (EMT) by MiR-34a-Mediated
Suppression of Notch-1 Expression in Non-Small Cell Lung Cancer Cell Lines. J. Biol. Chem. 2013, 288,
27343–27357. [CrossRef] [PubMed]
216. Lin, C.-H.; Shen, Y.-A.; Hung, P.-H.; Yu, Y.-B.; Chen, Y.-J. Epigallocathechin Gallate, Polyphenol Present in
Green Tea, Inhibits Stem-like Characteristics and Epithelial-Mesenchymal Transition in Nasopharyngeal
Cancer Cell Lines. BMC Complement. Altern. Med. 2012, 12, 201. [CrossRef] [PubMed]
217. Lin, Y.-S.; Tsai, P.-H.; Kandaswami, C.C.; Cheng, C.-H.; Ke, F.-C.; Lee, P.-P.; Hwang, J.-J.; Lee, M.-T. Effects of
Dietary Flavonoids, Luteolin, and Quercetin on the Reversal of Epithelial-Mesenchymal Transition in A431
Epidermal Cancer Cells. Cancer Sci. 2011, 102, 1829–1839. [CrossRef] [PubMed]
218. Hara, Y. Tea Catechins and Their Applications as Supplements and Pharmaceutics. Pharmacol. Res. 2011, 64,
100–104. [CrossRef] [PubMed]
© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
114
nutrients
Review
The Role of Vitamin E in Immunity
Ga Young Lee 1,2 and Sung Nim Han 1,2, *
1 Department of Food and Nutrition, College of Human Ecology, Seoul National University,
Seoul 08826, Korea; lgykiki90@snu.ac.kr
2 Research Institute of Human Ecology, Seoul National University, Seoul 08826, Korea
* Correspondence: snhan@snu.ac.kr; Tel.: +82-2-880-6836
Abstract: Vitamin E is a fat-soluble antioxidant that can protect the polyunsaturated fatty acids
(PUFAs) in the membrane from oxidation, regulate the production of reactive oxygen species (ROS)
and reactive nitrogen species (RNS), and modulate signal transduction. Immunomodulatory effects
of vitamin E have been observed in animal and human models under normal and disease conditions.
With advances in understating of the development, function, and regulation of dendritic cells (DCs),
macrophages, natural killer (NK) cells, T cells, and B cells, recent studies have focused on vitamin E’s
effects on specific immune cells. This review will summarize the immunological changes observed
with vitamin E intervention in animals and humans, and then describe the cell-specific effects of
vitamin E in order to understand the mechanisms of immunomodulation and implications of vitamin
E for immunological diseases.
1.2. Sources
The major dietary sources of vitamin E are vegetable oils. Nuts are good sources of vitamin
E as well [2]. Soybean, sunflower, corn, walnut, cottonseed, palm, and wheat germ oils contain
relatively higher amounts (more than approximately 50 mg vitamin E/100 g oil) of vitamin E than
other oils. The proportions of α-, β-, γ-, and δ-tocopherols vary depending on the oil type. Safflower
and sunflower oils are high in α-tocopherol, soybean and corn oils contain mainly γ-tocopherol,
and cottonseed oil contains similar proportions of α- and γ-tocopherols. Therefore, the types of oils
consumed through the diet affect the dietary intake levels of α-tocopherol. Vitamin E supplements are
quite popular and contribute considerably to vitamin E intake among some populations. Either natural
or synthetic forms of α-tocopherol are used as supplements.
Despite the relatively higher intake of γ-tocopherol from the diet than α-tocopherol, α-tocopherol
is the major form of vitamin E in the circulation because α-tocopherol transfer protein (α-TTP) has the
preferential binding affinity for α-tocopherol. α-TTP is involved in the transfer of α-tocopherol to the
plasma membrane [1].
1.3. Functions
Vitamin E is a major fat-soluble antioxidant that scavenges peroxyl radicals and terminates the
oxidation of polyunsaturated fatty acids (PUFAs). In the presence of vitamin E, peroxyl radicals react
with α-tocopherol instead of lipid hydroperoxide, the chain reaction of peroxyl radical production
is stopped, and further oxidation of PUFAs in the membrane is prevented [1]. Tocopheroxyl
radicals—produced from α-tocopherol and peroxyl radicals—are reduced by vitamin C or glutathione,
form tocopherol dimers, undergo further oxidation, or act as prooxidants. The antioxidant activity of
vitamin E may be responsible for the regulation of several enzymes involved in signal transduction
because the activity of signaling enzymes is regulated by the redox state.
Vitamin E inhibits protein kinase C (PKC) activity by increasing PKC-α dephosphorylation
through the activation of protein phosphatase 2A. The inhibition of PKC by vitamin E has been
reported in various cells, and consequently, the inhibition of platelet aggregation; reduced proliferation
of monocytes, macrophages, neutrophils, and vascular smooth muscle cells; and decreased superoxide
production in neutrophils and macrophages have been observed [3,4].
Vitamin E may directly bind to the enzymes involved in the generation of lipid mediators or to
the transport proteins involved in signal transduction. Vitamin E may affect the membrane protein
116
Nutrients 2018, 10, 1614
interaction and translocation of the enzymes to the plasma membrane and therefore change the activity
of signal transduction enzymes [4].
117
Table 1. Modulation of immune responses by vitamin E in animal models.
Pregnant cows (n = 24/group) of gestation to day 21 of NA ↑IgG and IgA concentration in sow plasma Wang et al. 2017 [9]
lactation
Domestic cats (39 castrated male
225, 450 mg/kg diet for
and 33 intact female) (n = α-tocopherol ↑Lymphocyte proliferation (ConA, PHA) O’ Brien et al. 2015 [10]
28 days
8/group)
Young and old mice 500 mg/kg diet for 6 ↑Lymphocyte proliferation in old (ConA, PHA)
DL -α-tocotrienol Ren et al. 2010 [11]
(n = 11–13/group) weeks ↑IL-1β production in young
50, 200 mg/kg diet for
Young rats (n = 6/group) ↑Lymphocyte proliferation (ConA, LPS) Bendich et al. 1986 [12]
8–10 weeks
↑Lymphocyte proliferation (ConA, LPS)
500 mg/kg diet for 6 ↑DTH response
Old mice (n = 10/group) DL -α-tocopheryl acetate Meydani et al. 1986 [13]
weeks ↑IL-2 production
↓PGE2 production
118
↑Lymphocyte proliferation (ConA) in young
Young and old mice (n = 5/group) 500 IU (500 mg) for 9 DL -α-tocopherol acetate ↔Lymphocyte proliferation (ConA) in old Wakikawa et al. 1999 [14]
weeks
↑IFN-γ in young under restraint stress
Form of Vitamin E
Subjects Age Amount and Duration of Supplementation Effects on Immune Function References
Used
Young (n = 5) and senior 4.6 ± 0.3 mg/100 mL of vitamin E-enriched
18–25, 35–57 α-tocopherol acetate ↑15LOX2, TNF-α expression Capo et al. 2016 [20]
athletes (n = 5) beverage 5 days/week for 5 weeks
D -α-tocotrienol
↑IL-4 (TT vaccine), IFN-γ (ConA)
D -γ-tocotrienol Mahalingam et al. 2011
Healthy women (n = 108) 400 mg TRF/day for 56 days
Nutrients 2018, 10, 1614
18–25
[21]
D -δ-tocotrienol
↓IL-6 (LPS)
D -α-tocopherol
119
Healthy elderly males and
≥65 60, 200, 800 mg/day for 235 days DL -α-tocopherol ↑DTH and antibody titer to hepatitis B with 200, 800 mg Meydani et al. 1997 [27]
females (n = 88)
↑No. of positive DTH reaction with 100 mg
Healthy elderly males and DL -α-tocopheryl ↑dDiameter of induration of DTH reaction in a
65–80 50, 100 mg/day for 6 months Pallast et al. 1999 [28]
females (n = 161) acetate subgroup supplemented with 100 mg
↔IL-2 production
↓IFN-γ production
Healthy young adults
600 mg/day for 3 months40 mg/kg body
(n = 31) and premature 24–31 ↓Chemiluminescence Okano et al. 1990 [29]
weight for 8–14 days
infants (n = 10)
Cigarette smoker (n = 60) 33 ± 4 900 IU/day for 6 weeks ↓Chemiluminescence Richards et al. 1990 [30]
Prevented fish-oil-induced suppression of ConA
Healthy males (n = 40) 24–57 200 mg/day for 4 months all-rac-α-tocopherol Kramer et al. 1991 [31]
mitogenesis
↑DTH (maximal diameter) in 100, 200, 400 mg groups
Healthy elderly (n = 40) >65 100, 200, or 400 mg/day for 3 months DL -α-tocopherol Wu et al. 2006 [32]
↑Lymphocyte proliferation (ConA) in 200 mg group
Sedentary young and ↓IL-6 secretion
22–29, 55–74 800 IU (727 mg)/day for 48 days DL -α-tocopherol Cannon et al. 1991 [33]
elderly males (n = 21) ↓Exercise-enhanced IL-1β secretion
ConA, concanavalin A; DTH, delayed type hypersensitivity; IFN-γ, interferon-γ; 15LOX2, 15-lipoxygenase-2; PGE2 , prostaglandin E2 ; PHA, phytohemagglutinin; TRF, tocotrienol-rich
fraction; TT vaccine, tetanous toxoid vaccine.
Table 3. Effects of vitamin E supplementation on infectious diseases in animal models.
Dose and Duration of Form of Vitamin Infection Organism and Results: Effects of Vitamin E
Subjects Age References
Supplementation E Used Route of Infection Supplementation
Mice BALB/c 100 mg/kg for 8 days before MRSA, inoculated onto Higher NK cytotoxicity
6 months δ-, γ-Tocotrienol Pierpaoli et al. 2017 [34]
(n = 3–6/group) MRSA-challenge superficial surgical wounds Higher IL-24 mRNA expression levels
male mice 2, 22–26 500 mg/kg for 4 weeks prior to D -α-tocopheryl Streptococcus pneumoniae,
months infection proinflammatory cytokines (TNF-, IL-6) Bou Ghanem et al. 2015 [35]
C57BL/6 acetate intra-tracheally injected
(n = 6/group) were reduced
3-fold reduction in the number of PMNs
No difference in serum IgG or peripheral
Worm-free lambs 5.3 IU (3.56 mg)/kg BW for H. contortus L3 larvae, route
28–32 weeks D -α-tocopherol mRNA expression of IL-4 or IFN-γ De Wolf et al. 2014 [36]
(n =10/group) 12 weeks NA
Lower PCV, FEC, and worm burden
120
22 months 500mg/kg diet for 6 weeks (H3N2) by nasal Lower viral titre Hayek et al. 1997 [7]
(n = 4–9) acetate
inoculation
Mice, C57BL/6 160 IU/L liquid diet for 4, 8, 12, all-rac-α-tocopheryl Murine LP-BM5 leukaemia Restored IL-2 and IFN-γ production by
5 weeks Wang et al. 1994 [38]
(n = 6) 16 weeks acetate retrovirus by IP injection splenocytes following infection
1400 or 2800 mg orally once per Serum from vitamin E-supplemented
Calves, Holstein DL -α-tocopheryl Bovine rhinotracheitis
1d week, 1400 mg injection once calves inhibited the replication of bovine Reddy et al. 1986 [39]
(n = 7) acetate virus, in vitro
per week for 12 weeks rhinotracheitis virus in vitro
Mice, Swiss DL -α-tocopheryl Diplococcus pneumoniae type
4 weeks 180 mg/kg diet for 4 weeks Higher survival Heinzerling et al. 1974a [5]
Webster (n = 10) acetate I by IP injection
25 or 250 mg/kg bw orally for Pseudomonas aeruginosa,
Mice, BALB/C DL -α-tocopheryl
NA 4 days, starting 2 days before subeschar injection to Lower mortality rate Fang et al. 1990 [40]
(n = 25) acetate
burn injury burned mice
Vitamin E
Mice, BALB/C 4000mg/kg diet for 2, 4, or Listeria monocytogenes by IP
3 weeks injectable No difference in resistance Watson & Petro 1982 [41]
(NA) 14 weeks injection
(aqueous)
Rats,
180 mg/kg diet + 6000 IU DL -α-tocopheryl Mycoplasma pulmonis by
Sprague-Dawley 3 weeks Higher resistance to infection Tvedten et al. 1973 [42]
vitamin A/kg diet for 6 weeks acetate aerosol
(n = 6)
1000 IU orally, 300 mg/kg diet DL -α-tocopheryl Chlamydia by intratracheal Faster recovery (higher food intake and
Lambs (n = 10) NA Stephens et al. 1979 [43]
for 23 days acetate inoculation weight gains)
Table 3. Cont.
Dose and Duration of Form of Vitamin Infection Organism and Results: Effects of Vitamin E
Subjects Age References
Supplementation E Used Route of Infection Supplementation
200 mg/pig per day for 59 days Improved weight gain and recovery rate
DL -α-tocopheryl Treponema hyodysenteriae,
Pigs (n = 6) NA before infection and 22 days No beneficial effect on appetite and Teige et al. 1982 [45]
after infection acetate oral
diarrhoea
121
300 mg/kg diet for 6 weeks,
Chicks, broiler DL -α-tocopheryl
1 day starting 3 weeks before first E. coli, post-thoracic air sac Lower mortality Tengerdy & Nockels 1975 [49]
(n = 10) acetate
infection
Chicks, Leghorn 300 mg/kg diet for 4 weeks DL -α-tocopheryl
1 day E. coli by IV injection Lower mortality Likoff et al. 1981 [50]
(n = 22) before infection acetate
Vitamin E;
100, 000 mg/t diet for 10 weeks,
Tompson-Hayward,
Pigs (n = 10) 6–8 weeks starting 2 weeks before E. coli by IM injection Higher serum Ab titre Ellis & Vorhies 1976 [51]
Minneapolis, MN,
infection
USA
Ab, antibody; FEC, fecal egg count; HSV, Herpes simplex virus; MRSA, IFN-γ, interferon-γ; IM, intramuscular; IV, intravenous; Methicillin-resistant Staphylococcus aureus; NK, natural
killer; PCV, packed cell volume; PMN, polymorphonuclear leukocyte, RANTES, regulated on activation, normal T cell expressed and secreted; TNF-α, tumor necrosis factor-α.
Table 4. Effects of vitamin E supplementation on infectious diseases in humans.
Dose and Duration of Form of Vitamin Infection Organism and Route of Results: Effects of Vitamin E
Subjects Age References
Supplementation E Used Infection Supplementation
69% Lower incidence of pneumonia
among subgroups including participants
who smoked 5–19 cigarettes per day at
DL -α-tocopheryl
Male smoker 50–69 50 mg/d for median of 6 years Natural incidence of pneumonia baseline and exercised at leisure time Hemila et al. 2016 [52]
Nutrients 2018, 10, 1614
acetate
14% Lower incidence of pneumonia
among subgroups including participants
who smoked ≥20 cigarettes per day at
baseline and did not exercise
30 mg during pregnancy
HIV-infected (multivitamin form with 20 mg Natural incidence of malaria after Lower incidence of presumptive clinical
pregnant 25.4 vitamins B1, 20 mg B2, 25 mg B6, 100 NA having received malaria malaria, but higher risk of any malaria Olofin et al. 2014 [53]
Tanzanian women mg niacin, 50 μg B12, 500 mg C, prophylaxis during pregnancy parasitemia
and 800 μg folic acid)
Reduced glutathione (GSH) and
Patients with glutathione peroxidase, which are
900 IU (604.03 mg for D - or 818.18
HCV-related 54–75 α-tocopherol Natural incidence of cirrhosis significantly lower in cirrhotic patients Marotta et al. 2007 [54]
mg for DL-)/day for 6 months
cirrhosis (p < 0.05), were comparably improved by
vitamin E regimens
945 IU (634.23 mg)/day for 6 months
Patients with No difference in median log plasma Groenbak et al. 2006
122
18–75 with 500 mg ascorbic acid and 200 μg D -α-tocopherol Natural incidence of HCV
chronic HCV HCV-RNA [55]
of selenium
Fewer numbers of subjects with all and
Nursing home Natural incidence of respiratory upper respiratory infections
>65 200 IU/day for 1 year DL -α-tocopherol Meydani et al. 2004 [56]
residents infections Lower incidence of common cold
No effect on lower respiratory infection
Natural incidence of common cold Lower incidence of common cold
Male smokers 50–69 50 mg/day during 4-year follow-up α-tocopherol Hemila et al. 2002 [57]
episodes Reduction was greatest among older city
dwellers who smoked fewer than 15
cigarettes per day
No overall effect on the incidence of
DL -α-tocopheryl
Male smokers 50–69 years 50 mg/day for median of 6.1 years Natural incidence of pneumonia pneumonia. Hemila et al. 2004 [58]
acetate
Lower incidence of pneumonia among
the subjects who had initiated smoking
at a later age (>21)
Natural incidence and severity of
Non-institutionalized α-tocopherol No effect on incidence and severity of
>60 years 200 mg/day for median of 441 days self-reported acute respiratory tract Graat et al. 2002 [59]
individuals acetate acute respiratory tract infections
infections
HCV, hepatitis C virus.
Nutrients 2018, 10, 1614
3.1. Macrophages
Macrophages, important effector cells in the innate immune response, serve as antigen presenting
cells (APC) and regulate NK cells and T cells by producing cytokines, reactive oxygen species (ROS),
reactive nitrogen species (RNS), and prostaglandins. Cytokines produced by T cells and other immune
cells can shift the macrophages into different populations with distinct physiologies [60].
The effects of vitamin E on prostaglandin (PG)E2 production by macrophages from the aged
have been suggested as one of the mechanisms by which vitamin E improves the age-associated
decrease in the T cell-mediated immune response [61]. In a co-culture experiment in which purified
T cells and macrophages from young and old mice were cultured together, T cells from young mice
showed suppressed proliferation and IL-2 secretion when cultured with macrophages from old mice.
When macrophages from old mice were pre-incubated with 10 μg/mL vitamin E for 4 h, co-cultures
of old macrophages and young T cells showed significant improvement in proliferation. Vitamin E
pre-incubation of old macrophages improved proliferation and IL-2 production in co-cultures of old
macrophages and old T cells [62]. Macrophages from old mice produced significantly higher levels of
PGE2 , which was due to higher cyclooxygenase (COX) activity. Macrophages from old mice expressed
higher levels of inducible COX2 protein and mRNA [63]. These increases in PGE2 synthesis and COX
activity were lowered by in vivo vitamin E supplementation [64]. Macrophages isolated from old
mice fed a diet containing 500 ppm vitamin E for 30 days produced lower amounts of PGE2 and
had lower COX activity than those from old mice fed a control diet containing 30 ppm vitamin E,
but the COX2 mRNA levels and protein expression of the control and supplemented groups did not
123
Nutrients 2018, 10, 1614
differ. Thus, vitamin E’s effect on COX activity seemed to be through post-translational mechanisms
rather than through its effect at transcriptome or translational levels. In a subsequent study, it was
shown that vitamin E reduced COX activity in macrophages from old mice by decreasing peroxynitrite
production [65]. The inhibition of COX activity by vitamin E in old mice disappeared specifically with
the addition of a nitric oxide (NO) donor in the presence of a superoxide to elevate peroxynitrite levels
in the macrophage culture. There is a complex interplay between the nitric oxide synthase (NOS)
and COX pathways and NO increases COX2 activity, which seems to be due to the NO preventing
self-deactivation of COX by the superoxide as NO interacts with the superoxide [66].
In vivo supplementation of vitamin E (1500 IU D -α-tocopheryl acetate/day for 16 weeks)
in allergic asthmatic patients prevented the suppression of alveolar macrophage nuclear factor
(erythroid-derived 2)-like 2 (NRF2) activity after allergen challenge [67]. This study presented the
possibility of vitamin E’s protective role in allergies and asthmas through regulation of macrophage
NRF2 activity, but, further studies are needed to confirm the findings because of the small number of
patients (nine mild non-smoking allergic asthmatics) and the lack of appropriate controls.
124
Nutrients 2018, 10, 1614
were treated in culture starting from day 2, cells were stimulated on day 5, and then the surface
phenotype was determined on day 6. The expression of human leukocyte antigen(HLA)-DR, CD40
CD80, and CD86 appeared to be increased with lower concentrations of α-tocopherol (<0.05 mM), but
the combination of vitamin E and C prevented DC activation, as the upregulation of surface markers
was not observed. DCs treated with 0.5 mM vitamin E and 10 mM vitamin C showed lower levels
of intracellular ROS and inhibition of the nuclear factor (NF)-κB, PKC, and p38 mitogen-activated
protein kinase (MAPK) pathways. When bone marrow-derived dendritic cells (BMDCs) from Balb/c
mice were treated with 500 μM of α-tocopherol for 2 h, upregulation of phosphorylated inhibitor of κB
(IκB) by lipopolysaccharide (LPS)-stimulation was suppressed. Vitamin E treatment for 24 h resulted
in a reduced number of CD11+ CD86+ cells and ROS-positive cells, lower production of IL-12p70
and TNF-α, and decreased transwell migration of BMDCs. These effects of vitamin E on BMDCs
were partly dependent on Klotho expression. Vitamin E treatment on BMDCs resulted in higher
Klotho transcript and protein levels, and silencing of Klotho by transfection of Klotho siRNA abolished
the inhibitory effects of vitamin E on IL-12p70 production, number of ROS-positive cells, and DC
migration [76]. Klotho is a membrane protein that has been shown to mediate calcium transport into
the cells; regulate intracellular signaling pathways such as p53/p21, cyclin adenosine monophosphate
(cAMP), PKC, and Wnt; and inhibit the NF-κB pathway [77]. Therefore, the upregulation of Klotho
by vitamin E could be one of the mechanisms by which vitamin E modulates NF-κB mediated DC
function and maturation. However, the level of α-tocopherol used for in vitro treatment (500 μM) was
high and, therefore, further research is needed to elucidate the physiological relevance of vitamin E
treatment on the expression of Klotho and its involvement in the modulation of DC function.
In vivo supplementation of α-tocopherol at 150, 250, and 500 mg/kg diet in allergic female mice
reduced the lung CD11b+ DCs and mRNA levels of IL-4, IL-33, thymic stromal lymphopoietin (TSLP),
eotaxin 1 (CCL11), and eotaxin 2 (CCL24) in allergen challenged pups. Furthermore, when BMDCs
from 10-day-old neonates born to a control female were treated with 80 μM α-tocopherol for 24 h,
the number of CD45+ CD11b+ CD11+ DCs and the number of CD45+ CD11b+ CD11c+ Ly6c− MHCII−
DCs were reduced. Maternal supplementation with α-tocopherol was effective in decreasing allergic
responses in offspring from allergic mothers by affecting the development of subsets of DCs that
are critical for allergic responses [78]. On the other hand, γ-tocopherol supplementation exerted an
opposite response in the same model. In vivo supplementation of γ-tocopherol at 250 mg/kg diet in
allergic female mice resulted in a higher number of lung eosinophils, a higher number of lung CD11c+
CD11b+ DCs, and higher levels of lung lavage CCL11 in the offspring [79].
Modulation of the immune response by vitamin E has been observed in animal and human
studies, and DCs play a critical role in bridging innate and adaptive immune systems and initiating
adaptive immune responses. Despite the importance of DCs’ role in adaptive immune responses
and in diseases such as autoimmune diseases, few studies have investigated the DC-specific effect of
vitamin E.
3.4. T Cells
The effects of vitamin E on immune cells have been studied the most with T cells.
The dysregulation of immune function occurs with aging and the most significant changes are observed
in T cells. Age-associated changes in T cells include, but are not limited to, (1) defects in T cell receptor
(TCR) signal transduction such as a decrease in linker for the activation of T cells (LAT) phosphorylation
by zeta chain of T cell receptor associated protein kinase 70 (ZAP-70), (2) decreased intracellular influx
of calcium following stimulation, (3) diminished synapse formation, (4) diminished activation of the
mitogen activated protein kinase (MAP kinase) pathway, (5) decreased nuclear factor of activated
T-cells (NFAT) binding activity, and (6) a shift of the T cell population toward memory T cells [80]. As a
result, diminished production of IL-2 and reduced proliferative capacity of naive T cells are observed
and impaired T cell functions contribute to increased susceptibility to infectious diseases and poor
response to immunization.
125
Nutrients 2018, 10, 1614
Vitamin E has been shown to increase the cell division and IL-2 producing capacity of naïve T
cells, increase the percentage of T cells capable of forming an effective immune synapse, and reverse
the age-associated defect in the phosphorylation of LAT in T cells from old animals [81–83].
In vitro pre-incubation with 46 μM vitamin E for 4 h increased proliferation and IL-2 production in
T cells purified from old mice stimulated with anti-CD3 and anti-CD28. Increased IL-2 production was
due to both an increase in the number of activation-induced IL-2+ cells and an increase in the level of
IL-2 accumulated per cell. Vitamin E specifically increased the naive T cells’ ability to progress through
the cell division cycle in old mice [81]. The gene expression profile of T cells isolated from young and
old mice fed a diet supplemented with 500 ppm vitamin E for four weeks provided evidence that
vitamin E influences cell cycle-related molecules at the gene expression level. Higher expression of
cell cycle-related genes Ccnb2, Cdc2, and Cdc6 was observed in stimulated T cells from old mice fed
the vitamin E-supplemented diet compared with those fed the control diet, which was not observed
in young mice [84]. Cyclin B2, encoded by Ccnb2, binds to cyclin-dependent kinase 1 (also known
as Cdc2) and regulates the events during both the G2 /M transition and progression through mitosis.
Cdc6 is a key regulator in the early steps of DNA replication, as the binding of Cdc6 to chromatin is
a necessary and universal step in the acquisition of replication competences [85]. These alterations
in the expression of cell cycle-related genes observed with vitamin E might contribute to vitamin E
improving the proliferative ability of old T cells.
Marko et al. [82] showed that pre-incubation of CD4+ T cells isolated from old T cells with 46 μM
vitamin E for 4 h increased the percentage of CD4+ T cells displaying effective immune synapses.
Redistribution of Zap70, LAT, Vav, and phospholipase Cγ (PLCγ) into immune synapse increased
significantly with vitamin E treatment. This change was confirmed with in vivo supplementation
of vitamin E. In old mice fed a diet containing 500 ppm vitamin E for eight weeks, LAT and Vav
showed significantly higher redistribution into the T cell/APC contact area when purified CD4+
T cells were stimulated with murine CD3ε hybridoma. In a subsequent study, it was shown that
vitamin E could reverse the age-associated defect in the phosphorylation of LAT on tyrosine 191 [83].
The phosphorylation of LAT is required for the recruitment of adaptor and effector proteins. Therefore,
it plays a pivotal role in the assembly of microcluster structures in the initiation of T cell activation
signals. This evidence suggests that vitamin E can modulate the early stages of T cell activation.
Vitamin E seems to modulate Th1 and Th2 responses. The polarization of CD4 T cells to T
helper (Th)1 or Th2 cells has implications for the protection against different pathogens (intracellular
vs. extracellular pathogens) and the development of different types of chronic diseases (inflammatory
vs. allergic diseases). PBMCs isolated from allergic donors treated with vitamin E (12.5–50 μM)
showed dose-dependent decreases in IL-4 production [86]. IL-4 mRNA levels in activated PBMCs
were downregulated by 25 μM vitamin E treatment. Jurkat T cells treated with 50 μM vitamin E
exhibited downregulation of IL-4 promoter activity, which might be related to vitamin E blocking the
interaction of transcription factors with PRE-1 and P1. In vivo supplementation of vitamin E enhancing
the Th1 response has been observed in mice infected with influenza virus and in colorectal cancer
patients [6,87]. In colorectal cancer patients, two weeks of supplementation with 750 mg vitamin E
led to an increased frequency of IL-2 producing CD4+ T cells and increased IFN-γ production [87].
In old mice infected with influenza virus, 500 ppm vitamin E supplementation for eight weeks prior
to infection lowered the viral titer in the lung, and this protective effect of vitamin E was associated
with the enhancement of Th1 response. IFN-γ production levels correlated negatively with viral
titer, and old mice fed a vitamin E-supplemented diet produced significantly higher levels of IFN-γ
and IL-2 [6]. The gene expression profile of T cells isolated from young and old mice fed a diet
supplemented with 500 ppm vitamin E for four weeks provided evidence that vitamin E influences the
Th1/Th2 balance at the gene expression level. The increase in IL-4 expression following stimulation
was lower in T cells from old mice fed the vitamin E-supplemented diet compared with those fed the
control diet, and the ratio of IFN-γ and IL-4 expression levels was significantly higher in the vitamin E
group than in the control group [84].
126
Nutrients 2018, 10, 1614
Vitamin E can affect activation-induced cell death in T cells. In vitro treatment of primary human
T cells with 25 μM vitamin E suppressed CD95L expression and activation-induced cell death [88].
The reduction of CD95L mRNA levels and the proportion of CD95L-positive cells were related to the
suppression of NF-κB and AP-1 binding to the CD95L promoter target site by vitamin E. On the other
hand, α-tocopheryl succinate was shown to trigger apoptosis in Jurkat cells with caspase-activation
involved [89].
3.5. B Cells
Vitamin E supplementation has been reported to enhance humoral responses. Higher antibody
responses have been observed in animals and humans [19,27]. However, it is hard to differentiate
whether vitamin E’s direct effect on B cells or indirect effect through T cells contributes to higher
antibody responses.
4. Conclusions
Vitamin E has been shown to enhance immune responses in animal and human models and
to confer protection against several infectious diseases. Suggested mechanisms involved with these
changes are (1) the reduction of PGE2 production by the inhibition of COX2 activity mediated through
decreasing NO production, (2) the improvement of effective immune synapse formation in naive T
cells and the initiation of T cell activation signals, and (3) the modulation of Th1/Th2 balance. Higher
NK activity and changes in dendritic function such as lower IL-12 production and migration were
observed with vitamin E, but underlying mechanisms need to be further elucidated
Several considerations are warranted for the advancement in our understanding of vitamin E’s
role in immunity. For in vitro studies to support implications for the regulation of immunological
diseases, the physiological relevance of vitamin E levels used for treatment should be considered.
Different forms of vitamin E exert differential effects on immune cells. Cell-specific effects of vitamin
E provide valuable evidence regarding the immunomodulatory mechanisms of vitamin E, but the
interplay between immune cells should not be ignored, because interactions between immune cells are
critical in the regulation of immune function.
Author Contributions: Literature search and manuscript preparation were performed by G.Y.L. and S.N.H.
The manuscript was revised and finalized by S.N.H.
Funding: This work was supported by the Basic Science Research Program through the National Research
Foundation of Korea (NRF) funded by the Ministry of Education (grant number NRF-2018R1D1A1B07049178).
Conflicts of Interest: The author declares no conflicts of interests.
References
1. Traber, M.G. Vitamin E regulatory mechanisms. Annu. Rev. Nutr. 2007, 27, 347–362. [CrossRef] [PubMed]
2. Sheppard, A.J.; Pennington, J.A.T.; Weihrauch, J.L. Analysis and distribution of vitamin E in vegetable oils
and foods. In Vitamin E in Health and Disease; Packer, L., Fuchs, J., Eds.; Marcel Dekker: New York, NY, USA,
1980; pp. 7–65.
3. Traber, M.G.; Atkinson, J. Vitamin E, antioxidant and nothing m more. Free Rad. Biol. Med. 2007, 43, 4–15.
[CrossRef] [PubMed]
4. Zingg, J.M. Vitamin E: A role in signal transduction. Annu. Rev. Nutr. 2015, 35, 135–173. [CrossRef] [PubMed]
5. Heinzerling, R.H.; Tengerdy, R.P.; Wick, L.L.; Lueker, D.C. Vitamin E protects mice against Diplococcus
pneumoniae type I infection. Infect. Immun. 1974, 10, 1292–1295. [PubMed]
6. Han, S.N.; Wu, D.; Ha, W.K.; Beharka, A.; Smith, D.E.; Bender, B.S.; Meydani, S.N. Vitamin E supplementation
increases T helper 1 cytokine production in old mice infected with influenza virus. Immunology 2000, 100,
487–493. [CrossRef] [PubMed]
7. Hayek, M.G.; Taylor, S.F.; Bender, B.S.; Han, S.N.; Meydani, M.; Smith, D.E.; Eghtesada, S.; Meydani, S.N.
Vitamin E supplementation decreases lung virus titers in mice infected with influenza. J. Infect. Dis. 1997,
176, 273–276. [CrossRef] [PubMed]
127
Nutrients 2018, 10, 1614
8. Dalia, A.M.; Loh, T.C.; Sazili, A.Q.; Jahromi, M.F.; Samsudin, A.A. Effects of vitamin E, inorganic selenium,
bacterial organic selenium, and their combinations on immunity response in broiler chickens. BMC Vet. Res.
2018, 14, 249. [CrossRef] [PubMed]
9. Wang, L.; Xu, X.; Su, G.; Shi, B.; Shan, A. High concentration of vitamin E supplementation in sow diet during
the last week of gestation and lactation affects the immunological variables and antioxidative parameters in
piglets. J. Dairy Res. 2017, 84, 8–13. [CrossRef] [PubMed]
10. O’Brien, T.; Thomas, D.G.; Morel, P.C.; Rutherfurd-Markwick, K.J. Moderate dietary supplementation with
vitamin E enhances lymphocyte functionality in the adult cat. Res. Vet. Sci. 2015, 99, 63–69. [CrossRef]
[PubMed]
11. Ren, Z.; Pae, M.; Dao, M.C.; Smith, D.; Meydani, S.N.; Wu, D. Dietary supplementation with tocotrienols
enhances immune function in C57BL/6 mice. J. Nutr. 2010, 140, 1335–1341. [CrossRef] [PubMed]
12. Bendich, A.; Gabriel, E.; Machlin, L.J. Dietary vitamin E requirement for optimum immune responses in the
rat. J. Nutr. 1986, 116, 675–681. [CrossRef] [PubMed]
13. Meydani, S.N.; Meydani, M.; Verdon, C.P.; Shapiro, A.A.; Blumberg, J.B.; Hayes, K.C. Vitamin E
supplementation suppresses prostaglandin E1(2) synthesis and enhances the immune response of aged mice.
Mech. Ageing Dev. 1986, 34, 191–201. [CrossRef]
14. Wakikawa, A.; Utsuyama, M.; Wakabayashi, A.; Kitagawa, M.; Hirokawa, K. Vitamin E enhances the immune
functions of young but not old mice under restraint stress. Exp. Gerontol. 1999, 34, 853–862. [CrossRef]
15. Moriguchi, S.; Kobayashi, N.; Kishino, Y. High dietary intakes of vitamin E and cellular immune functions in
rats. J. Nutr. 1990, 120, 1096–1102. [CrossRef] [PubMed]
16. Sakai, S.; Moriguchi, S. Long-term feeding of high vitamin E diet improves the decreased mitogen response
of rat splenic lymphocytes with aging. J. Nutr. Sci. Vitaminol. 1997, 43, 113–122. [CrossRef] [PubMed]
17. Reddy, P.G.; Morrill, J.L.; Minocha, H.C.; Stevenson, J.S. Vitamin E is Immunostimulatory in calves.
J. Dairy Sci. 1987, 70, 993–999. [CrossRef]
18. Tanaka, J.; Fujiwara, H.; Torisu, M. Vitamin E and immune response. I. Enhancement of helper T cell activity
by dietary supplementation of vitamin E in mice. Immunology 1979, 38, 727–734. [PubMed]
19. Beharka, A.A.; Han, S.N.; Adolfsson, O.; Wu, D.; Lipman, R.; Smith, D.; Cao, G.; Meydani, M.; Meydani, S.N.
Long-term dietary antioxidant supplementation reduces production of selected inflammatory mediators by
murine macrophages. Nutr. Res. 2000, 20, 281–296. [CrossRef]
20. Capó, X.; Martorell, M.; Sureda, A.; Riera, J.; Drobnic, F.; Tur, J.A.; Pons, A. Effects of Almond- and Olive
Oil-Based Docosahexaenoic- and Vitamin E-Enriched Beverage Dietary Supplementation on Inflammation
Associated to Exercise and Age. Nutrients. 2016, 8, 619. [CrossRef] [PubMed]
21. Mahalingam, D.; Radhakrishnan, A.K.; Amom, Z.; Ibrahim, N.; Nesaretnam, K. Effects of supplementation
with tocotrienol-rich fraction on immune response to tetanus toxoid immunization in normal healthy
volunteers. Eur. J. Clin. Nutr. 2011, 65, 63–69. [CrossRef] [PubMed]
22. Radhakrishnan, A.K.; Lee, A.L.; Wong, P.F.; Kaur, J.; Aung, H.; Nesaretnam, K. Daily supplementation of
tocotrienol-rich fraction or alpha-tocopherol did not induce immunomodulatory changes in healthy human
volunteers. Br. J. Nutr. 2009, 101, 810–815. [CrossRef] [PubMed]
23. Prasad, J.S. Effect of vitamin E supplementation on leukocyte function. Am. J. Clin. Nutr. 1980, 33, 606–608.
[CrossRef] [PubMed]
24. Harman, D.; Miller, R.W. Effect of vitamin E on the immune response to influenza virus vaccine and the
incidence of infectious disease in man. Age 1986, 9, 21–23. [CrossRef]
25. Meydani, S.N.; Barklund, M.P.; Liu, S.; Meydani, M.; Miller, R.A.; Cannon, J.G.; Morrow, F.D.; Rocklin, R.;
Blumberg, J.B. Vitamin E supplementation enhances cell-mediated immunity in healthy elderly subjects.
Am. J. Clin. Nutr. 1990, 52, 557–563. [CrossRef] [PubMed]
26. De Waart, F.G.; Portengen, L.; Doekes, G.; Verwaal, C.J.; Kok, F.J. Effect of 3 months vitamin E
supplementation on indices of the cellular and humoral immune response in elderly subjects. Br. J. Nutr.
1997, 78, 761–774. [CrossRef] [PubMed]
27. Meydani, S.N.; Meydani, M.; Blumberg, J.B.; Leka, L.S.; Siber, G.; Loszewski, R.; Thompson, C.; Pedrosa, M.C.;
Diamond, R.D.; Stollar, B.D. Vitamin E supplementation and in vivo immune response in healthy elderly
subjects. A randomized controlled trial. JAMA 1997, 277, 1380–1386. [CrossRef] [PubMed]
128
Nutrients 2018, 10, 1614
28. Pallast, E.G.; Schouten, E.G.; de Waart, F.G.; Fonk, H.C.; Doekes, G.; von Blomberg, B.M.; Kok, F.J. Effect of
50- and 100-mg vitamin E supplements on cellular immune function in noninstitutionalized elderly persons.
Am. J. Clin. Nutr. 1999, 69, 1273–1281. [CrossRef] [PubMed]
29. Okano, T.; Tamai, H.; Mino, M. Superoxide generation in leukocytes and vitamin E. Int. J. Vitam. Nutr. Res.
1991, 61, 20–26. [PubMed]
30. Richards, G.A.; Theron, A.J.; van Rensburg, C.E.; van Rensburg, A.J.; van der Merwe, C.A.; Kuyl, J.M.;
Anderson, R. Investigation of the effects of oral administration of vitamin E and beta-carotene on the
chemiluminescence responses and the frequency of sister chromatid exchanges in circulating leukocytes
from cigarette smokers. Am. Rev. Respir. Dis. 1990, 142, 648–654. [CrossRef] [PubMed]
31. Kramer, T.R.; Schoene, N.; Douglass, L.W.; Judd, J.T.; Ballard-Barbash, R.; Taylor, P.R.; Bhagavan, H.N.;
Nair, P.P. Increased vitamin E intake restores fish-oil-induced suppressed blastogenesis of mitogen-stimulated
T lymphocytes. Am. J. Clin. Nutr. 1991, 54, 896–902. [CrossRef] [PubMed]
32. Wu, D.; Han, S.N.; Meydani, M.; Meydani, S.N. Effect of concomitant consumption of fish oil and vitamin
E on T cell mediated function in the elderly: A randomized double-blind trial. J. Am. Coll. Nutr. 2006, 25,
300–306. [CrossRef] [PubMed]
33. Cannon, J.G.; Meydani, S.N.; Fielding, R.A.; Fiatarone, M.A.; Meydani, M.; Farhangmehr, M.; Orencole, S.F.;
Blumberg, J.B.; Evans, W.J. Acute phase response in exercise. II. Associations between vitamin E, cytokines,
and muscle proteolysis. Am. J. Physiol. 1991, 260, 1235–1240. [CrossRef] [PubMed]
34. Pierpaoli, E.; Orlando, F.; Cirioni, O.; Simonetti, O.; Giacometti, A.; Provinciali, M. Supplementation with
tocotrienols from Bixa orellana improves the in vivo efficacy of daptomycin against methicillin-resistant
Staphylococcus aureus in a mouse model of infected wound. Phytomedicine 2017, 36, 50–53. [CrossRef]
[PubMed]
35. Bou Ghanem, E.N.; Clark, S.; Du, X.; Wu, D.; Camilli, A.; Leong, J.M.; Meydani, S.N. The α-tocopherol
form of vitamin E reverses age-associated susceptibility to streptococcus pneumoniae lung infection by
modulating pulmonary neutrophil recruitment. J. Immunol. 2015, 194, 1090–1099. [CrossRef] [PubMed]
36. De Wolf, B.M.; Zajac, A.M.; Hoffer, K.A.; Sartini, B.L.; Bowdridge, S.; LaRoith, T.; Petersson, K.H. The effect
of vitamin E supplementation on an experimental Haemonchus contortus infection in lambs. Vet. Parasitol.
2014, 205, 140–149. [CrossRef] [PubMed]
37. Sheridan, P.A.; Beck, M.A. The immune response to herpes simplex virus encephalitis in mice is modulated
by dietary vitamin E. J. Nutr. 2008, 138, 130–137. [CrossRef] [PubMed]
38. Wang, Y.; Huang, D.S.; Eskelson, C.D.; Watson, R.R. Long-Term Dietary Vitamin E Retards Development of
Retrovirus-Induced Disregulation in Cytokine Production. Clin. Immunol. Immunopathol. 1994, 72, 70–75.
[CrossRef] [PubMed]
39. Reddy, P.G.; Morrill, J.L.; Minocha, H.C.; Morrill, M.B.; Dayton, A.D.; Frey, R.A. Effect of supplemental
vitamin E on the immune system of calves. J. Dairy Sci. 1986, 69, 164–171. [CrossRef]
40. Fang, C.H.; Peck, M.D.; Alexander, J.W.; Babcock, G.F.; Warden, G.D. The effect of free radical scavengers on
outcome after infection in burned mice. J. Trauma. 1990, 30, 453–456. [CrossRef] [PubMed]
41. Watson, R.; Petro, T.M. Cellular immune response, corticosteroid levels and resistance to Listeria
monocytogenes and murine leukemia in mice fed a high vitamin E diet. N.Y. Acad. Sci. 1982, 393, 205–210.
[CrossRef]
42. Tvedten, H.W.; Whitehair, C.K.; Langham, R.F. Influence of vitamins A and E on gnotobiotic and
conventionally maintained rats exposed to Mycoplasma pulmonis. J. Am. Vet. Med. Assoc. 1973, 163,
605–612. [PubMed]
43. Stephens, L.C.; McChesney, A.E.; Nockels, C.F. Improved recovery of vitamin E-treated lambs that have been
experimentally infected with intratracheal Chlamydia. Br. Vet. J. 1979, 135, 291–293. [CrossRef]
44. Schildknecht, E.G.; Squibb, R.L. The effect of vitamins A, E and K on experimentally induced histomoniasis
in turkeys. Parasitology 1979, 78, 19–31. [CrossRef] [PubMed]
45. Teige, J.; Tollersrud, S.; Lund, A.; Larsen, H.J. Swine dysentery: The influence of dietary vitamin E and
selenium on the clinical and pathological effects of Treponema hyodysenteriae infection in pigs. Res. Vet. Sci.
1982, 32, 95–100. [PubMed]
46. Tengerdy, R.P.; Meyer, D.L.; Lauerman, L.H.; Lueker, D.C.; Nockels, C.F. Vitamin E-enhanced humoral
antibody response to Clostridium perfringens type D in sheep. Br. Vet. J. 1983, 139, 147–152. [CrossRef]
129
Nutrients 2018, 10, 1614
47. Smith, K.L.; Harrison, J.H.; Hancock, D.D.; Todhunter, D.A.; Conrad, H.R. Effect of vitamin E and selenium
supplementation on incidence of clinical mastitis and duration of clinical symptoms. J. Dairy Sci. 1984, 67,
1293–1300. [CrossRef]
48. Heinzerling, R.H.; Nockels, C.F.; Quarles, C.L.; Tengerdy, R.P. Protection of chicks against E.coli infection by
dietary supplementation with vitamin E. Proc. Soc. Exp. Biol. Med. 1974, 146, 279–283. [CrossRef] [PubMed]
49. Tengerdy, R.P.; Nockels, C.F. Vitamin E or vitamin A protects chickens against E. coli infection. Poult. Sci.
1975, 54, 1292–1296. [CrossRef] [PubMed]
50. Likoff, R.O.; Guptill, D.R.; Lawrence, L.M.; McKay, C.C.; Mathias, M.M.; Nockels, C.F.; Tengerdy, R.P. Vitamin
E and aspirin depress prostaglandins in protection of chickens against Escherichia coli infection. Am. J.
Clin. Nutr. 1981, 34, 245–251. [CrossRef] [PubMed]
51. Ellis, R.P.; Vorhies, M.W. Effect of supplemental dietary vitamin E on the serologic response of swine to an
Escherichia coli bacterin. J. Am. Vet. Med. Assoc. 1976, 168, 231–232. [PubMed]
52. Hemilä, H. Vitamin E administration may decrease the incidence of pneumonia in elderly males.
Clin. Interv. Aging 2016, 11, 1379–1385. [CrossRef] [PubMed]
53. Olofin, I.O.; Spiegelman, D.; Aboud, S.; Duggan, C.; Danaei, G.; Fawzi, W.W. Supplementation with
multivitamins and vitamin A and incidence of malaria among HIV-infected Tanzanian women. J. Acquir.
Immune Defic. Syndr. 2014, 67 (Suppl. S4), S173–S178. [CrossRef] [PubMed]
54. Marotta, F.; Yoshida, C.; Barreto, R.; Naito, Y.; Packer, L. Oxidative-inflammatory damage in cirrhosis: Effect
of vitamin E and a fermented papaya preparation. J. Gastroenterol. Hepatol. 2007, 22, 697–703. [CrossRef]
[PubMed]
55. Groenbaek, K.; Friis, H.; Hansen, M.; Ring-Larsen, H.; Krarup, H.B. The effect of antioxidant supplementation
on hepatitis C viral load, transaminases and oxidative status: A randomized trial among chronic hepatitis C
virus-infected patients. Eur. J. Gastroenterol. Hepatol. 2006, 18, 985–989. [CrossRef] [PubMed]
56. Meydani, S.N.; Leka, L.S.; Fine, B.C.; Dallal, G.E.; Keusch, G.T.; Singh, M.F.; Hamer, D.H. Vitamin E and
respiratory tract infections in elderly nursing home residents: A randomized controlled trial. JAMA 2004,
292, 828–836. [CrossRef] [PubMed]
57. Hemilä, H.; Kaprio, J.; Albanes, D.; Heinonen, O.P.; Virtamo, J. Vitamin C, vitamin E, and beta-carotene in
relation to common cold incidence in male smokers. Epidemiology 2002, 13, 32–37. [CrossRef] [PubMed]
58. Hemilä, H.; Virtamo, J.; Albanes, D.; Kaprio, J. Vitamin E and beta-carotene supplementation and
hospital-treated pneumonia incidence in male smokers. Chest 2004, 125, 557–565. [CrossRef] [PubMed]
59. Graat, J.M.; Schouten, E.G.; Kok, F.J. Effect of daily vitamin E and multivitamin-mineral supplementation on
acute respiratory tract infections in elderly persons: A randomized controlled trial. JAMA 2002, 288, 715–721.
[CrossRef] [PubMed]
60. Mosser, D.M.; Edwards, J.P. Exploring the full spectrum of macrophage activation. Nat. Rev. Immunol. 2008,
8, 958–969. [CrossRef] [PubMed]
61. Meydani, S.N.; Han, S.N.; Wu, D. Vitamin E and immune response in the aged: Molecular mechanisms and
clinical implications. Immunol. Rev. 2005, 205, 269–284. [CrossRef] [PubMed]
62. Beharka, A.A.; We, D.; Han, S.N.; Meydani, S.N. Macrophage prostaglandin production contributes to
the age-associated decrease in T cell function which is reversed by the dietary antioxidant vitamin E.
Mech. Age. Dev. 1997, 93, 59–77. [CrossRef]
63. Hayek, M.G.; Mura, C.; Wu, D.; Beharka, A.A.; Han, S.N.; Paulson, K.E.; Hwang, D.; Meydani, S.N. Enhanced
expression of inducible cyclooxygenase with age in murine macrophages. J. Immunol. 1997, 159, 2445–2451.
[PubMed]
64. Wu, D.; Mura, C.; Beharka, A.A.; Han, S.N.; Paulson, K.E.; Hwang, D.; Meydani, S.N. Age-associated increase
in PGE2 synthesis and COX activity in murine macrophages is reversed by vitamin E. Am. J. Physiol. 1998,
275, C661–C668. [CrossRef] [PubMed]
65. Beharka, A.A.; Wu, D.; Serafini, M.; Meydani, S.N. Mechanism of vitamin E inhibition of cyclooxygenase
activity in macrophages from old mice: Role of peroxynitrite. Free. Radic. Biol. Med. 2002, 32, 503–511.
[CrossRef]
66. Sorokin, A. Nitric Oxide Synthase and Cyclooxygenase Pathways: A Complex Interplay in Cellular Signaling.
Curr. Med. Chem. 2016, 23, 2559–2578. [CrossRef] [PubMed]
130
Nutrients 2018, 10, 1614
67. Dworski, R.; Han, W.; Blackwell, T.S.; Hoskins, A.; Freeman, M.L. Vitamin E prevents NRF2 suppression
by allergens in asthmatic alveolar macrophages in vivo. Free Radic. Biol. Med. 2011, 51, 516–521. [CrossRef]
[PubMed]
68. Adachi, N.; Migita, M.; Ohta, T.; Higashi, A.; Matsuda, I. Depressed natural killer cell activity due to
decreased natural killer cell population in a vitamin E-deficient patient with Shwachman syndrome:
Reversible natural killer cell abnormality by alpha-tocopherol supplementation. Eur. J. Pediatr. 1997,
156, 444–448. [CrossRef] [PubMed]
69. Ravaglia, G.; Forti, P.; Maioli, F.; Bastagli, L.; Facchini, A.; Mariani, E.; Savarino, L.; Sassi, S.; Cucinotta, D.;
Lenaz, G. Effect of micronutrient status on natural killer cell immune function in healthy free-living subjects
aged ≥90 y. Am. J. Clin. Nutr. 2000, 71, 590–598. [CrossRef] [PubMed]
70. Hanson, M.G.; Ozenci, V.; Carlsten, M.C.; Glimelius, B.L.; Frödin, J.E.; Masucci, G.; Malmberg, K.J.;
Kiessling, R.V. A short-term dietary supplementation with high doses of vitamin E increases NK cell
cytolytic activity in advanced colorectal cancer patients. Cancer Immunol. Immunother. 2007, 56, 973–984.
[CrossRef] [PubMed]
71. Stiff, A.; Trikha, P.; Mundy-Bosse, B.; McMichael, E.; Mace, T.A.; Benner, B.; Kendra, K.; Campbell, A.;
Gautam, S.; Abood, D.; et al. Nitric Oxide Production by Myeloid-Derived Suppressor Cells Plays a Role
in Impairing Fc Receptor-Mediated Natural Killer Cell Function. Clin. Cancer Res. 2018, 24, 1891–1904.
[CrossRef] [PubMed]
72. Ganguly, D.; Haak, S.; Sisirak, V.; Reizis, B. The role of dendritic cells in autoimmunity. Nat. Rev. Immunol.
2013, 13, 566–577. [CrossRef] [PubMed]
73. Pearce, E.J.; Everts, B. Dendritic cell metabolism. Nat. Rev. Immunol. 2015, 15, 18–29. [CrossRef] [PubMed]
74. Alloatti, A.; Kotsias, F.; Magalhaes, J.G.; Amigorena, S. Dendritic cell maturation and cross-presentation:
Timing matters! Immunol. Rev. 2016, 272, 97–108. [CrossRef] [PubMed]
75. Tan, P.H.; Sagoo, P.; Chan, C.; Yates, J.B.; Campbell, J.; Beutelspacher, S.C.; Foxwell, B.M.; Lombardi, G.;
George, A.J. Inhibition of NF-kappa B and oxidative pathways in human dendritic cells by antioxidative
vitamins generates regulatory T cells. J. Immunol. 2005, 174, 7633–7644. [CrossRef] [PubMed]
76. Xuan, N.T.; Trang, P.T.; Van Phong, N.; Toan, N.L.; Trung, D.M.; Bac, N.D.; Nguyen, V.L.; Hoang, N.H.; van
Hai, N. Klotho sensitive regulation of dendritic cell functions by vitamin E. Biol. Res. 2016, 49, 45. [CrossRef]
[PubMed]
77. Buendía, P.; Ramírez, R.; Aljama, P.; Carracedo, J. Klotho Prevents Translocation of NFκB. Vitam. Horm. 2016,
101, 119–150. [PubMed]
78. Abdala-Valencia, H.; Berdnikovs, S.; Soveg, F.W.; Cook-Mills, J.M. α-Tocopherol supplementation of allergic
female mice inhibits development of CD11c+CD11b+ dendritic cells in utero and allergic inflammation in
neonates. Am. J. Physiol. Lung Cell. Mol. Physiol. 2014, 307, L482–L496. [CrossRef] [PubMed]
79. Abdala-Valencia, H.; Soveg, F.; Cook-Mills, J.M. γ-Tocopherol supplementation of allergic female mice
augments development of CD11c+CD11b+ dendritic cells in utero and allergic inflammation in neonates.
Am. J. Physiol. Lung Cell. Mol. Physiol. 2016, 310, L759–L771. [CrossRef] [PubMed]
80. Molano, A.; Meydani, S.N. Vitamin E, signalosomes and gene expression in T cells. Mol. Aspects. Med. 2012,
33, 55–62. [CrossRef] [PubMed]
81. Adolfsson, O.; Huber, B.T.; Meydani, S.N. Vitamin E-enhanced IL-2 production in old mice: Naive but not
memory T cells show increased cell division cycling and IL-2-producing capacity. J. Immunol. 2001, 167,
3809–3817. [CrossRef] [PubMed]
82. Marko, M.G.; Ahmed, T.; Bunnell, S.C.; Wu, D.; Chung, H.; Huber, B.T.; Meydani, S.N. Age-associated
decline in effective immune synapse formation of CD4(+) T cells is reversed by vitamin E supplementation.
J. Immunol. 2007, 178, 1443–1449. [CrossRef] [PubMed]
83. Marko, M.G.; Pang, H.J.; Ren, Z.; Azzi, A.; Huber, B.T.; Bunnell, S.C.; Meydani, S.N. Vitamin E reverses
impaired linker for activation of T cells activation in T cells from aged C57BL/6 mice. J. Nutr. 2009, 139,
1192–1197. [CrossRef] [PubMed]
84. Han, S.N.; Adolfsson, O.; Lee, C.K.; Prolla, T.A.; Ordovas, J.; Meydani, S.N. Age and vitamin E-induced
changes in gene expression profiles of T cells. J. Immunol. 2006, 177, 6052–6061. [CrossRef] [PubMed]
85. Pelizon, C. Down to the origin: Cdc6 protein and the competence to replicate. Trends Cell Biol. 2003, 13,
110–113. [CrossRef]
131
Nutrients 2018, 10, 1614
86. Li-Weber, M.; Giaisi, M.; Treiber, M.K.; Krammer, P.H. Vitamin E inhibits IL-4 gene expression in peripheral
blood T cells. Eur. J. Immunol. 2002, 32, 2401–2408. [CrossRef]
87. Malmberg, K.J.; Lenkei, R.; Petersson, M.; Ohlum, T.; Ichihara, F.; Glimelius, B.; Frödin, J.E.; Masucci, G. A
short-term dietary supplementation of high doses of vitamin E increases T helper 1 cytokine production in
patients with advanced colorectal cancer. Clin. Cancer Res. 2002, 8, 1772–1778. [PubMed]
88. Li-Weber, M.; Weigand, M.A.; Giaisi, M.; Süss, D.; Treiber, M.K.; Baumann, S.; Ritsou, E.; Breitkreutz, R.;
Krammer, P.H. Vitamin E inhibits CD95 ligand expression and protects T cells from activation-induced cell
death. J. Clin. Invest. 2002, 110, 681–690. [CrossRef] [PubMed]
89. Neuzil, J.; Svensson, I.; Weber, T.; Weber, C.; Brunk, U.T. α-tocopheryl succinate-induced apoptosis in Jurkat
T cells involves caspase-3 activation, and both lysosomal and mitochondrial destabilisation. FEBS Lett. 1999,
445, 295–300. [CrossRef]
© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
132
nutrients
Review
The Microbiotic Highway to Health— New
Perspective on Food Structure, Gut Microbiota,
and Host Inflammation
Nina Wærling Hansen 1 and Anette Sams 2, *
1 Molecular Endocrinology Unit (KMEB), Department of Endocrinology, Institute of Clinical Research,
University of Southern Denmark, DK-5000 Odense, Denmark; nwhansen@health.sdu.dk
2 Department of Clinical Experimental Research, Glostrup Research Institute, Copenhagen University
Hospital, Nordstjernevej 42, DK-2600 Glostrup, Denmark
* Correspondence: anette.nielsen.03@regionh.dk
Abstract: This review provides evidence that not only the content of nutrients but indeed the
structural organization of nutrients is a major determinant of human health. The gut microbiota
provides nutrients for the host by digesting food structures otherwise indigestible by human enzymes,
thereby simultaneously harvesting energy and delivering nutrients and metabolites for the nutritional
and biological benefit of the host. Microbiota-derived nutrients, metabolites, and antigens promote the
development and function of the host immune system both directly by activating cells of the adaptive
and innate immune system and indirectly by sustaining release of monosaccharides, stimulating
intestinal receptors and secreting gut hormones. Multiple indirect microbiota-dependent biological
responses contribute to glucose homeostasis, which prevents hyperglycemia-induced inflammatory
conditions. The composition and function of the gut microbiota vary between individuals and
whereas dietary habits influence the gut microbiota, the gut microbiota influences both the nutritional
and biological homeostasis of the host. A healthy gut microbiota requires the presence of beneficial
microbiotic species as well as vital food structures to ensure appropriate feeding of the microbiota.
This review focuses on the impact of plant-based food structures, the “fiber-encapsulated nutrient
formulation”, and on the direct and indirect mechanisms by which the gut microbiota participate in
host immune function.
1. Introduction
The gut microbiota is a complex ecosystem residing in the gastro-intestinal (GI) tract and consists
of a diverse microbiotic community living in symbiosis with the host. A diverse microbiota is
considered a major positive regulator of the interdependent metabolic and immune function of the
host. The gut microbiota is shaped alongside the host immune system in a synergistic partnership [1].
Trillions of microorganisms participate in the maturation and regulation of immune function, energy
metabolism, and hormonal balance [2], including regulation of intestinal mucosal barriers, fermentation
of undigested nutrients, and synthesis of short chain fatty acids (SCFA) [3]. Furthermore, the microbiota
releases microbiota-derived antigens [4,5], vitamins [6,7], and other molecules, such as tryptophan
metabolites [8,9], that interact with the host biology.
The dynamic relationship between the microbiotic ecosystem and its host is emphasized by
the joint utilization of consumed nutrients. Diet is a shared substrate between the host and the gut
microbiota and the choice of diet affects host health both directly and indirectly by affecting the
abundance and composition of the microbiotic community [10,11].
The structural organization of complex carbohydrates determines the site of digestion and
absorption in the GI tract [12], and hence the level of cross feeding [13], community organization,
and proliferation of specific microorganisms. Importantly, all plant cell walls consist of complex
carbohydrate structures that are digested selectively by microbiotic enzymes and not by host enzymes.
Thus, complex polysaccharides in plant cell walls function as both substrate for the microbiota and as
a sustained release delivery system of other plant-cell derived nutrients and biomolecules for both
microbiotic and host utilization. The structural properties of the cell wall vary between different plant
cells [14], and in theory determine the specific site of plant cell wall degradation and nutrient release.
Plant-based foods can be placed on a moving scale from whole foods (unrefined) to fully refined
foods, and currently no official distinction based on the degrees of refinement exists. Foods with added
refined nutrients (e.g., sugar, starch or oils) or any food, which has been structurally disrupted on the
macroscopic level are considered refined. The degree of refinement is thus determined by the degree of
macroscopic structural disruption (e.g., intact whole grain vs. ground whole grain vs. sifted flour) and
the degree of refined supplementation. When plant-based food structures are completely disrupted
requiring no microbiotic interaction for digestion we consider the food fully refined.
Microbiota composition varies greatly among different cultures and individuals [15,16],
and specific genera of bacteria are more abundant in healthy individuals as compared to individuals
in different disease states, as shown, for example, for Bifidobacterium and Lactobacillus [17–19], whose
functions include carbohydrate fermentation and vitamin synthesis [6,20,21]. The degradation and
utilization of complex carbohydrates rely on specific microbiotic enzymes secreted from the microbiota
in the lower intestine [22]. The fermentation process yields energy for microbiotic proliferation
and metabolites, e.g. SCFA [23] for regulation of inflammatory responses [24] and gut hormone
secretion [25] in the host. The latter is of great importance for glucose homeostasis and plays a major
role for the metabolic and inflammatory health of the host [26].
Microbiota composition is also closely connected to the integrity of our intestinal mucosal barrier.
Upon microbiota dysbiosis mucosal bacteria can impair epithelial function and cause increased gut
permeability with consequent immune dysregulation, leading to inflammatory disease states [27].
A number of disorders with an inflammatory component including obesity, inflammatory bowel
diseases (IBD), type 2 diabetes (T2D), colorectal cancer, and cardiovascular diseases have been linked
with dysbiosis [28–32]. This association presents the gut microbiota as a potentially modifiable factor
in the etiology of these conditions. It is therefore of great importance for public and individual health
to recognize that microbiota composition may be diversified within days or weeks upon introduction
of unrefined plant-based diets [16].
In clinical trials diets are often classified as low carbohydrate vs. high carbohydrate, and low
fiber vs. high fiber. These trials lack discrimination between refined and unrefined carbohydrates.
Acknowledging the degree of refinement would likely strengthen the reproducibility of diet
intervention studies for reasons discussed in the present review [33–35].
This paper analyzes the interplay between structures of plant-based foods or “fiber-encapsulated
nutrients”, the gut microbiota, and their joint impact on host nutrient bioavailability, immune function,
and metabolic function, and proposes a simple tool to select plant-based foods from a nutritional-
and microbiota-promoting perspective. The review provides a biological explanation of the health
superiority of structurally maintained and unrefined foods compared to nutrient-matched refined
foods, combining classic and current scientific knowledge within nutrition, pharmaceutical science,
microbiology, and plant biology. In addition, we add the perspective of “fiber as nutrient encapsulation”
to the well-established molecular benefits of fiber as microbiotic nutrient (prebiotic). This structural
perspective should be included in the current nutritional guidelines, which focus on nutrient and fiber
content rather than nutrient and fiber structure.
134
Nutrients 2018, 10, 1590
135
Nutrients 2018, 10, 1590
Figure 1. Schematic illustration of an animal and a plant cell. The cell wall (mainly complex
carbohydrates), vacuoles (usually starch or lipid storage, depending on cell type), and chloroplasts
are plant-cell specific structures that serve as nutrients for the microbiota, whereas the cellular content
serves as signal molecules for host metabolic and immune regulation as well as nutrients for both
host and microbiota. An intact plant cell is defined as unrefined food whereas the degree of structural
disruption (and the supplementation of additives) defines the degree of refinement.
The most characteristic component found in all plant cell walls is the fiber, cellulose [43], which
consists of a collection of β-1,4-linked glucan chains that interact with each other via hydrogen
bonds to form a crystalline microfibril [44]. In addition to cellulose, the primary plant cell wall
contains two categories of matrix polysaccharides-pectins and hemicelluloses, and several proteins
and glycoproteins, including various enzymes and structural proteins [45].
Intracellular plant compartments and their relative ratio constitute the macro- and micronutrients
in the plant and the dispersion of macronutrients varies from plant to plant, changing during aging
with a larger proportion of protein and lipids during early growth [46]. In addition to macronutrients
and vitamins, the encapsulated plant cells also contain phytonutrients, such as polyphenols (flavonoids
and stilbenes), carotenoids, plant sterols, and poly-unsaturated fatty acids (PUFA), many of which
are used as nutraceutical ingredients and exhibit beneficial effects on metabolic and cardiovascular
parameters [47].
Although monomers from the cell wall degradation process are absorbed via host transporters,
most of the plant-cell wall functions as microbiota substrate [12]. Furthermore, the mixture of
macronutrients, vitamins, and phytonutrients encapsulated within the plant cell membrane have
a two-tier function, with part of their host-health-promoting effects also benefiting the microbiota.
In other words, the microbiota and the host share both host-enzyme-resistant carbohydrates and
fiber-encapsulated macro- and micronutrients.
4. Microbiota Composition
In human adults, five bacterial phyla have been reported to be dominant in the gut microbiota
of healthy individuals: Firmicutes, Bacteroidetes, Actinobacteria, Proteobacteria and Verrucomicrobia [48],
with more than 90% of the species belonging to Firmicutes and Bacteroidetes. Representatives of the
other phyla comprise from 2 to 10% with great interpersonal variation [49]. Changes in the relative
abundance of the two dominant bacterial divisions, Bacteroidetes and Firmicutes, have been associated
with an increased BMI [50]. An overrepresentation of butyrate-producing species, such as Roseburia spp.
and Eubacterium spp. within the Firmicutes phylum affects the metabolic potential of the microbiota,
i.e., the microbiotic capacity to harvest energy from the diet and produce SCFA available for host
consumption [51].
In humans, an abundance and diversity in polysaccharide-fermenting bacteria, such as species
within the Bacteroidetes phylum, are inversely related to obesity and other metabolic disorders
with an inflammatory component [52–55]. Members of Bacteroidetes express very large numbers of
genes that encode microbiotic enzymes and readily can switch between different energy sources
in the gut depending on substrate availability. Within the Bacteroidetes phyla, two genera are
136
Nutrients 2018, 10, 1590
particularly dominant and mutual antagonistic; Bacteroides and Prevotella [56]. The relative ratio
between these two genera varies, with the Bacteroides-driven type reported as dominant in individuals
with a higher intake of protein and animal fat, and Prevotella more common in individuals who consume
more carbohydrate-rich diets [57,58] These findings represent distinct community composition
types—so-called enterotypes—based on genus composition [59], suggesting that such compositional
differences both reflect dietary intake and determine an individual’s response to different diets [55].
Long-term dietary habits have considerable effect on the gut microbiota as evidence by differences
in microbiotic compositions across geographical and/or cultural-dependent diets [28,57,60]. Microbiota
composition also respond to short-term macronutrient changes; moreover, functional traits linking
enterotypes to diet have been identified [57,58]. Animal and human studies show that an acute dietary
switch results in dramatic shifts in the gut microbiota within the course of 24 h [16]. An animal-based
diet increases the abundance of bile-tolerant genera (Alistipes, Bilophila, and Bacteroides) and decreases
the level of Firmicutes species capable of fermenting dietary plant polysaccharides. Microbiotic activity
mirrors differences between herbivorous and carnivorous mammals, reflecting trade-offs between
carbohydrate and protein fermentation [61]. Analyses of the relative abundance of taxonomic groups
have shown that short-term animal-based diets have a greater impact on microbiotic community
structure compared to plant-based diets [16], suggesting perhaps a default microbiotic setting favoring
plant-based diets.
As mentioned above, bacterial species have varying abilities to degrade fibers from different
sources. The ability to degrade a wider variety of complex carbohydrates therefore forms a more
diverse bacterial community. Moreover, the biodiversity, function, and abundance of a microbiotic
community also depend on cross-feeding among members, which is the process where essential
metabolites synthesized by a subset of the community provide substrate for growth of another/other
bacterial species [62]. These microbiotic interactions create a complex pattern of interconnected
metabolisms and link members of different genera/phyla together in communities.
Biodiversity therefore contributes to the stability and resilience of the bacterial communities.
A high level of resilience ensures renewal and reorganization of the community if a species is
lost, whereas communities with low biodiversity are more susceptible to loss of individual species.
Highly diverse ecosystems most likely contain species with redundant functions, making the loss of
a single species tolerable, because other functionally redundant species can take over [63]. Evidently,
the microbiotic composition and community organization are reflective of dietary habits and affect
host metabolic and immune processes. The distinction between enterotypes with distinct fermentation
kinetics should therefore be considered an environmental factor that contributes to both health
and disease.
137
Nutrients 2018, 10, 1590
Throughout the GI tract the level of glucose transporters is highly regulated by hexose availability
and insulin levels [70–73]. In addition, the expression of glucose transporters has been shown to be
increased in e.g., obesity and diabetes [74,75] and as a result, certain individuals experience a larger
bioavailability of monosaccharides.
Despite individual differences in host digestive enzyme- and nutrient transporter expression,
a general notion must be highlighted: The density and diversity of the microbiota increases with the
distance from the stomach. The more complex the carbohydrate, the longer the journey until complete
fermentation, and therefore the more likely that carbohydrates are utilized by the microbiota and not
by the host.
138
Nutrients 2018, 10, 1590
Figure 2. Illustration of interactions between plant cells, microbiota, and the host epithelium. Plant cells
are encapsulated by the fibrous cell wall, which is a microbiota-selective nutrient. The cell wall also
serves as a transport vehicle for the additional plant cell contents as well as a vehicle for nutrients to be
shared between the microbiota and the host upon release from its plant cell structure. The microbiota
secretes glycosidases, peptidases, and lipases, thereby releasing molecular nutrients for transport into
epithelial or bacterial cells.
If plant cells have been refined, this organization does not remain intact, and even if the fibers
reach the microbiota (e.g., fiber supplements), only limited amounts of amino acids and lipids will
reach the microbes.
Some bacterial genera e.g., Bacteroides [92] and Lactobacillus [93] secrete both peptidases and
glycosidases and the more distal fiber-fermenting species Bacteroides thetaiotaomicron also express
dipeptidylpeptidases (DPPs) [94,95]. Interestingly, the DPP4 antagonist vildagliptin, used for the
treatment of T2D, induces a higher relative abundance of Bacteroidetes, a lower abundance of Firmicutes,
and thus a reduced ratio of Firmicutes/Bacteroidetes in rats [96]. This is one among many examples of
anti-diabetic drugs that has recently been shown to induce secondary health promoting actions via the
microbiota [97].
Another pharmaceutical exploration of the microbiota is the study of bacterial peptide transporters
with the aim to identify new antibiotic targets in virulent bacteria. Since blockage of peptide
transporters is a potent antibiotic pathway, specific bacteria are dependent on their endogenous
peptide transporters and on amino acid fuel [98]. In both gram negative and gram positive virulent
bacteria, two distinct vital types of peptide transporters exist, proton-dependent peptide transporters
(PRT or POT) [99] and ATP-binding cassettes (ABC transporters) [100].
As in the case of amino acids, energy dependent fatty acid transporters are also present in both
outer and inner membranes of bacterial strains. Even Lactobacillus expresses amino acid and fatty
139
Nutrients 2018, 10, 1590
acid transporters [83], which supports the bacterial ability and need to consume fatty acids slowly
released from plant cells during the microbiotic digestion of the plant cell wall. In specific bacteria,
the process requires the outer membrane-bound fatty acid transport protein FadL and the inner
membrane associated fatty acyl CoA synthetase (FACS) [101,102].
The diversity in fiber-coated plant cell structures also serves as delivery system for local acting
biologically active substances, e.g., taste receptor agonist for distal enterocytes and the microbiota [103],
and maturation agents for the immune system [104] which will be discussed below.
140
Nutrients 2018, 10, 1590
Imbalances in immunological homeostasis can occur due to the loss of habitat-specific species and
symbionts, resulting in an invasion of opportunistic species not normally able to colonize that specific
habitat within the intestinal mucosa. Such a mismatch between species and habitat triggers a potentially
pathogenic host response [130]. The detection of pathogens by the host is achieved through the families
of PRR localized in the intestinal epithelium. PRR recognize conserved molecular structures known as
microbe-associated molecular patterns and induce production of innate effector molecules [131,132].
The most well-known PRR are the toll-like receptors (TLR), RAGE, and the NOD-like receptors
expressed on the intestinal epithelial cells. Activation of TLR by bacterial products promotes epithelial
cell proliferation, secretion of IgA into the gut lumen, and expression of antimicrobial peptides,
thereby establishing a microorganism-induced system of epithelial cell homeostasis and repair in the
intestine [132].
Bacterial translocation, e.g., infiltration of lipopolysaccharide (LPS) through the intestinal barrier,
can promote local and systemic immune responses via PRR. Monocolonization of GF mice by
Escherichia coli is, for example, sufficient to induce macrophage infiltration, and polarization towards
pro-inflammatory phenotype of immune cells in the adipose tissue [133], contributing to a state of
low-grade inflammation in the adipose tissue. Disruption of intestinal barrier integrity by viable
bacteria has been attributed to various intestinal inflammatory diseases such as IBD, celiac disease,
irritable bowel syndrome, and colorectal cancer [32] as well as to diseases in extra-intestinal organs,
such as the liver [134].
The major microbiota-derived SCFA, butyrate, acetate, and propionate are implicated in multiple
anti-inflammatory mechanisms [135]. Besides being used by colonocytes as a source of ATP, SCFA can
act as extracellular signaling molecules that activate cell-surface free fatty acids receptors (FFAR)
(G-protein-coupled receptors (GPCR)) or inhibit histone deacetylases (HDAC) [136]. FFAR are
differentially expressed on adipocytes, immune cells, and enteroendocrine L-cells. Depending on the
location of the receptor and the amount and type of SCFA, the response can have multiple, variable
downstream effects [3]. FFAR, expressed on immune cells, such as macrophages, dendritic cells,
and neutrophils will upon activation by butyrate inhibit the release of pro-inflammatory cytokines, or
stimulate differentiation of T regulatory (Treg) cells [136].
Activation of FFAR2/GPR43 by butyrate on enteroendocrine L-cells of the ileum and colon induces
production and secretion of intestinal peptide YY (PYY), glucagon-like peptide 1 (GLP-1) [25,137].
An abundant production of SCFA can therefore indirectly regulate blood glucose levels via
the insulinotropic effect of GLP-1 and induce satiety by the anorexigenic effect of PYY on the
hypothalamus [138]. In addition, enteroendocrine cells in the duodenum are dependent on
microbe-mediated mechanisms to express cholecystokinin (CKK), a gut peptide hormone responsible
for stimulating fat absorption. GF mice have exhibited a reduced number of these cells with impact on
their fat absorption [139].
SCFA can also act by inhibition of histone deacetylase (HDAC) activity. HDAC is related to
a suppression of malignant transformation and a stimulation of apoptosis of precancerous colonic
cells [140] as well as to the stimulation of epithelial production of retinoic acid (RA) [141]. RA is
involved in many physiological processes, including regulation of IgA [142] and the polarization of
specialized mucosal myeloid cells [143].
Specific phytonutrients entail several immunomodulatory functions. Some are mediated by a shift
in microbiotic composition favoring the abundance of specific bacteria with health promoting effects
or create occupancy resistance to enteric pathogens with pro-inflammatory potential [144]. Others are
receptor-mediated, e.g., genistein, a flavonoid compound present in legumes, which can positively
effect β-cell mass and mitigate T2D in mice [145,146].
In summary, the biodiversity of the microbiotic community is critical for both inter-microbiotic
interactions and host-microbe engagements. In a healthy and proper fed microbiota the inflammatory
and immunoregulatory species are in balance and the diverse bacterial communities act together to
produce metabolites with direct effect on host health.
141
Nutrients 2018, 10, 1590
Figure 3. Schematic overview of interactions between food, microbiota, and host. Most of a refined diet
bypasses the microbiota, providing a high nutrient bioavailability for the host. In contrast, unrefined,
plant-based food structures provide a lower nutrient bioavailability for the host, using the difference
to develop and maintain a healthy microbiota. The healthy microbiota provides short chain fatty
acids (SCFA), antioxidants, and vitamins for the host along with direct immunomodulatory effects.
The high monosaccharide bioavailability in refined foods introduces a risk of hyperglycemia, which
cause numerous direct and indirect pro-inflammatory actions [26] (not shown in figure).
142
Nutrients 2018, 10, 1590
growing prevalence of non-communicable diseases, insulin resistance, overweight, and obesity [147],
it is obvious that current nutritional guidelines need to be revisited to re-establish the endogenous
microbiota-induced health promotion. The current nutritional guidelines are mainly based on optimal
molecular nutrient content for host metabolism, while establishment of synergies between food, host,
and microbiotic community are absent.
Nutritional guidelines need to consider both the content and the structure of carbohydrates,
because the microbiotic breakdown of carbohydrates strengthens our microbiotic ally in support of
host health. Furthermore, it should also reflect that consumption of “fiber-formulated nutrients”
is essential to maintain a healthy diverse microbiota to balance our nutritional, immunological,
and metabolic health.
We have developed a simple tool to visualize healthy microbiota-engaging food choices (Figure 4).
This tool divides foods into four categories depending on their carbohydrate content and the degree of
refinement. Healthy, unrefined foods with low relative carbohydrate content (green box) should make
up the plant-based bulk of an adult basic diet. These foods interact with the microbiota. Refined foods
or foods with high relative carbohydrate content (yellow boxes) should be considered carbohydrate or
energy supplementation, while the “comfort” foods in the red box, which are both refined and high in
carbohydrate content, should be used sparingly considering their potentially negative health impact
and their lack of microbiotic health promotion.
143
Nutrients 2018, 10, 1590
11. Conclusions
The increasing prevalence of non-communicable diseases with an inflammatory component has
led to the understanding of the gut microbiota as an intrinsic regulator of host immune responses.
There is a growing awareness of the role of microbiotic communities, microbiota-derived products,
and more particularly of the link between these components and disease states in humans. The ability
to target immune pathways relies on the understanding of both acute and long-term impacts on the
gut microbiota. The microbiotic communities can adapt to host diet in multiple ways by the interaction
among different bacterial species, loss or proliferation of specific species, and by interaction with the
host. Diet therefore presents itself as a microbiota-modulating factor and hence a health-modulating
factor. A revision of the current nutritional guidelines considering both the molecular content and
molecular structure of nutrients for insulin resistant and insulin sensitive individuals, could be valuable
for the restoration of beneficial bacteria and microbiota diversity, enabling a shift from disease to health
promoting states in the individual as well as in the general population.
Author Contributions: Conceptualization, A.S.; Writing-Original Draft Preparation, N.W.H and A.S;
Writing-Review & Editing, N.W.H. and A.S.; Visualization, N.W.H and A.S.
Funding: This research received no external funding.
Acknowledgments: We are grateful to Anne Stæhr Johansen, PhD, for her editorial assistance.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Francino, M.P. Birth Mode-Related Differences in Gut Microbiota Colonization and Immune System
Development. Ann. Nutr. Metab. 2018, 73, 12–16. [CrossRef] [PubMed]
2. Gill, S.R.; Pop, M.; DeBoy, R.T.; Eckburg, P.B.; Turnbaugh, P.J.; Samuel, B.S.; Gordon, J.I.; Relman, D.A.;
Fraser-Liggett, C.M.; Nelson, K.E. Metagenomic Analysis of the Human Distal Gut Microbiome. Science
2006, 312, 1355–1359. [CrossRef] [PubMed]
3. Morrison, D.J.; Preston, T. Formation of short chain fatty acids by the gut microbiota and their impact on
human metabolism. Gut Microbes 2016, 7, 189–200. [CrossRef] [PubMed]
4. Ivanov, I.I.; Atarashi, K.; Manel, N.; Brodie, E.L.; Shima, T.; Karaoz, U.; Wei, D.; Goldfarb, K.C.; Santee, C.A.;
Lynch, S.V.; et al. Induction of Intestinal Th17 Cells by Segmented Filamentous Bacteria. Cell 2009, 139,
485–498. [CrossRef] [PubMed]
5. Ivanov, I.I.; Frutos Rde, L.; Manel, N.; Yoshinaga, K.; Rifkin, D.B.; Sartor, R.B.; Finlay, B.B.; Littman, D.R.
Specific microbiota direct the differentiation of IL-17-producing T-helper cells in the mucosa of the small
intestine. Cell. Host Microbe 2008, 4, 337–349. [CrossRef] [PubMed]
6. Rossi, M.; Amaretti, A.; Raimondi, S. Folate Production by Probiotic Bacteria. Nutrients 2011, 3, 118–134.
[CrossRef] [PubMed]
7. LeBlanc, J.G.; Milani, C.; de Giori, G.S.; Sesma, F.; van Sinderen, D.; Ventura, M. Bacteria as vitamin suppliers
to their host: A gut microbiota perspective. Curr. Opin. Biotechnol. 2013, 24, 160–168. [CrossRef] [PubMed]
8. Shimada, Y.; Kinoshita, M.; Harada, K.; Mizutani, M.; Masahata, K.; Kayama, H.; Takeda, K. Commensal
bacteria-dependent indole production enhances epithelial barrier function in the colon. PLoS ONE 2013,
8, e80604. [CrossRef] [PubMed]
9. Bansal, T.; Alaniz, R.C.; Wood, T.K.; Jayaraman, A. The bacterial signal indole increases epithelial-cell
tight-junction resistance and attenuates indicators of inflammation. Proc. Natl. Acad. Sci. USA 2010, 107,
228–233. [CrossRef] [PubMed]
10. Cotillard, A.; Kennedy, S.P.; Kong, L.C.; Prifti, E.; Pons, N.; Le Chatelier, E.; Almeida, M.; Quinquis, B.;
Levenez, F.; Galleron, N. Dietary intervention impact on gut microbial gene richness. Nature 2013, 500,
585–588. [CrossRef] [PubMed]
11. Le Chatelier, E.; Nielsen, T.; Qin, J.J.; Prifti, E.; Hildebrand, F.; Falony, G.; Almeida, M.; Arumugam, M.;
Batto, J.M.; Kennedy, S. Richness of human gut microbiome correlates with metabolic markers. Nature 2013,
500, 541–546. [CrossRef] [PubMed]
144
Nutrients 2018, 10, 1590
12. Koropatkin, N.M.; Cameron, E.A.; Martens, E.C. How glycan metabolism shapes the human gut microbiota.
Nat. Rev. Microbiol. 2012, 10, 323–335. [CrossRef] [PubMed]
13. Falony, G.; De Vuyst, L. Prebiotics and Probiotics Science and Technology; Springer: New York, NY, USA, 2009;
pp. 639–679.
14. Keegstra, K. Plant cell walls. Plant. Physiol. 2010, 154, 483–486. [CrossRef] [PubMed]
15. De Filippo, C.; Cavalieri, D.; Di Paola, M.; Ramazzotti, M.; Poullet, J.B.; Massart, S.; Collini, S.; Pieraccini, G.;
Lionetti, P. Impact of diet in shaping gut microbiota revealed by a comparative study in children from Europe
and rural Africa. Proc. Natl. Acad. Sci. USA 2010, 107, 14691–14696. [CrossRef] [PubMed]
16. David, L.A.; Maurice, C.F.; Carmody, R.N.; Gootenberg, D.B.; Button, J.E.; Wolfe, B.E.; Ling, A.V.; Devlin, A.S.;
Varma, Y.; Fischbach, M.A. Diet rapidly and reproducibly alters the human gut microbiome. Nature 2014,
505, 559–563. [CrossRef] [PubMed]
17. Clemente, J.C.; Ursell, L.K.; Parfrey, L.W.; Knight, R. The Impact of the Gut Microbiota on Human Health:
An Integrative View. Cell 2012, 148, 1258–1270. [CrossRef] [PubMed]
18. Tojo, R.; Suárez, A.; Clemente, M.G.; de los Reyes-Gavilán, C.G.; Margolles, A.; Gueimonde, M.;
Ruas-Madiedo, P. Intestinal microbiota in health and disease: Role of bifidobacteria in gut homeostasis.
World J. Gastroenterol. 2014, 20, 15163–15176. [CrossRef] [PubMed]
19. Marchesi, J.R.; Adams, D.H.; Fava, F.; Hermes, G.D.A.; Hirschfield, G.M.; Hold, G.; Quraishi, M.N.; Kinross, J.;
Smidt, H.; Tuohy, K.M. The gut microbiota and host health: A new clinical frontier. Gut 2016, 65, 330–339.
[CrossRef] [PubMed]
20. Bottacini, F.; Ventura, M.; Sinderen, D.; Motherway, M. Diversity, ecology and intestinal function of
bifidobacteria. Microb. Cell. Fact. 2014, 13, S4. [CrossRef] [PubMed]
21. Wells, J.M. Immunomodulatory mechanisms of lactobacilli. Microb. Cell. Fact. 2011, 10, S17. [CrossRef]
[PubMed]
22. El Kaoutari, A.; Armougom, F.; Gordon, J.I.; Raoult, D.; Henrissat, B. The abundance and variety of
carbohydrate-active enzymes in the human gut microbiota. Nat. Rev. Microbiol. 2013, 11, 497–504. [CrossRef]
[PubMed]
23. Bach Knudsen, K.E. Microbial Degradation of Whole-Grain Complex Carbohydrates and Impact on
Short-Chain Fatty Acids and Health. Adv. Nutr. 2015, 6, 206–213. [CrossRef] [PubMed]
24. Maslowski, K.M.; Vieira, A.T.; Ng, A.; Kranich, J.; Sierro, F.; Yu, D.; Schilter, H.C.; Rolph, M.S.; Mackay, F.;
Artis, D. Regulation of inflammatory responses by gut microbiota and chemoattractant receptor GPR43.
Nature 2009, 461, 1282–1286. [CrossRef] [PubMed]
25. Tolhurst, G.; Heffron, H.; Lam, Y.S.; Parker, H.E.; Habib, A.M.; Diakogiannaki, E.; Cameron, J.; Grosse, J.;
Reimann, F.; Gribble, F.M. Short-chain fatty acids stimulate glucagon-like peptide-1 secretion via the
G-protein-coupled receptor FFAR2. Diabetes 2012, 61, 364–371. [CrossRef] [PubMed]
26. Hansen, N.W.; Hansen, A.J.; Sams, A. The endothelial border to health: Mechanistic evidence of the
hyperglycemic culprit of inflammatory disease acceleration. IUBMB Life 2017, 69, 148–161. [CrossRef]
[PubMed]
27. Bischoff, S.C.; Barbara, G.; Buurman, W.; Ockhuizen, T.; Schulzke, J.D.; Serino, M.; Tilg, H.; Watson, A.;
Wells, J.M. Intestinal permeability—A new target for disease prevention and therapy. BMC Gastroenterol.
2014, 14, 189. [CrossRef] [PubMed]
28. Ley, R.E.; Turnbaugh, P.J.; Klein, S.; Gordon, J.I. Microbial ecology: Human gut microbes associated with
obesity. Nature 2006, 444, 1022–1023. [CrossRef] [PubMed]
29. Ley, R.E.; Bäckhed, F.; Turnbaugh, P.; Lozupone, C.A.; Knight, R.D.; Gordon, J.I. Obesity alters gut microbial
ecology. Proc. Natl. Acad. Sci. USA 2005, 102, 11070–11075. [CrossRef] [PubMed]
30. Tang, W.H.W.; Kitai, T.; Hazen, S.L. Gut Microbiota in Cardiovascular Health and Disease. Circ. Res. 2017,
120, 1183–1196. [CrossRef] [PubMed]
31. Serino, M.; Blasco-Baque, V.; Nicolas, S.; Burcelin, R. Managing the manager: Gut microbes, stem cells and
metabolism. Diabetes Metab. 2014, 40, 186–190. [CrossRef] [PubMed]
32. Kang, M.; Martin, A. Microbiome and colorectal cancer: Unraveling host-microbiota interactions in
colitis-associated colorectal cancer development. Semin. Immunol. 2017, 32, 3–13. [CrossRef] [PubMed]
145
Nutrients 2018, 10, 1590
33. Gardner, C.D.; Trepanowski, J.F.; Del Gobbo, L.C.; Hauser, M.E.; Rigdon, J.; Ioannidis, J.P.A.; Desai, M.;
King, A.C. Effect of Low-Fat vs Low-Carbohydrate Diet on 12-Month Weight Loss in Overweight Adults
and the Association With Genotype Pattern or Insulin Secretion: The DIETFITS Randomized Clinical Trial.
JAMA 2018, 319, 667–679. [CrossRef] [PubMed]
34. Hjorth, M.F.; Ritz, C.; Blaak, E.E.; Saris, W.H.; Langin, D.; Poulsen, S.K.; Larsen, T.M.; Sørensen, T.I.; Zohar, Y.;
Astrup, A. Pretreatment fasting plasma glucose and insulin modify dietary weight loss success: Results from
3 randomized clinical trials. Am. J. Clin. Nutr. 2017, 106, 499–505. [CrossRef] [PubMed]
35. Brouns, F. Overweight and diabetes prevention: Is a low-carbohydrate-high-fat diet recommendable?
Eur. J. Nutr. 2018, 57, 1301–1312. [CrossRef] [PubMed]
36. Dhital, S.; Warren, F.J.; Butterworth, P.J.; Ellis, P.R.; Gidley, M.J. Mechanisms of starch digestion by
α-amylase—Structural basis for kinetic properties. Crit. Rev. Food Sci. Nutr. 2017, 57, 875–892. [CrossRef]
[PubMed]
37. Englyst, H.N.; Kingman, S.M.; Cummings, J.H. Classification and measurement of nutritionally important
starch fractions. Eur. J. Clin. Nutr. 1992, 46, S33–S50. [PubMed]
38. Fry, S.C. Primary Cell Wall Metabolism: Tracking the Careers of Wall Polymers in Living Plant.
Source New Phytol. 2004, 161, 641–675. [CrossRef]
39. Preiss, J. Plant starch synthesis. In Starch Food, 2nd ed.; Woodhead Publishing: Sawston, Cambridge, UK,
2018; pp. 3–95.
40. Roberfroid, M.B. Introducing inulin-type fructans. Br. J. Nutr. 2005, 93, S13–S25. [CrossRef] [PubMed]
41. Gibson, G.R.; Scott, K.P.; Rastall, R.A.; Tuohy, K.M.; Hotchkiss, A.; Dubert-Ferrandon, A.; Gareau, M.;
Murphy, E.F.; Saulnier, D.; Loh, G. Dietary prebiotics: Current status and new definition. Food Sci. Technol.
Bull. Funct. Foods 2010, 7, 1–19. [CrossRef]
42. Scott, K.P.; Martin, J.C.; Duncan, S.H.; Flint, H.J. Prebiotic stimulation of human colonic butyrate-producing
bacteria and bifidobacteria, in vitro. FEMS Microbiol. Ecol. 2014, 87, 30–40. [CrossRef] [PubMed]
43. Alberts, B.; Johnson, A.; Lewis, J.; Raff, M.; Roberts, K.; Walter, P. The Plant Cell Wall. In Molecular Biology of
the Cell, 4th ed.; Garland Science: New York, NY, USA, 2002.
44. Somerville, C. Cellulose Synthesis in Higher Plants. Annu. Rev. Cell. Dev. Biol. 2006, 22, 53–78. [CrossRef]
[PubMed]
45. Rose, J.K.C.; Lee, S.J. Straying off the highway: Trafficking of secreted plant proteins and complexity in the
plant cell wall proteome. Plant. Physiol. 2010, 153, 433–436. [CrossRef] [PubMed]
46. Weber, N.; Taylor, D.C.; Underhill, E.W. Biosynthesis of storage lipids in plant cell and embryo cultures.
Adv. Biochem. Eng. Biotechnol. 1992, 45, 99–131. [PubMed]
47. Bagchi, M.; Patel, S.; Zafra-Stone, S.; Bagchi, D. Selected herbal supplements and nutraceuticals.
In Reproductive and Developmental Toxicology; Academic Press: Cambridge, UK, 2011; pp. 385–393.
48. Tremaroli, V.; Bäckhed, F. Functional interactions between the gut microbiota and host metabolism. Nature
2012, 489, 242–249. [CrossRef] [PubMed]
49. Turnbaugh, P.J.; Hamady, M.; Yatsunenko, T.; Cantarel, B.L.; Duncan, A.; Ley, R.E.; Sogin, M.L.; Jones, W.J.;
Roe, B.A.; Affourtit, J.P. A core gut microbiome in obese and lean twins. Nature 2009, 457, 480–484. [CrossRef]
[PubMed]
50. Koliada, A.; Syzenko, G.; Moseiko, V.; Budovska, L.; Puchkov, K.; Perederiy, V.; Gavalko, Y.; Dorofeyev, A.;
Romanenko, M.; Tkach, S. Association between body mass index and Firmicutes/Bacteroidetes ratio in
an adult Ukrainian population. BMC Microbiol. 2017, 17, 120. [CrossRef] [PubMed]
51. Louis, P.; Young, P.; Holtrop, G.; Flint, H.J. Diversity of human colonic butyrate-producing bacteria revealed
by analysis of the butyryl-CoA:acetate CoA-transferase gene. Environ. Microbiol. 2010, 12, 304–314. [CrossRef]
[PubMed]
52. InterAct Consortium. Dietary fibre and incidence of type 2 diabetes in eight European countries: The
EPIC-InterAct Study and a meta-analysis of prospective studies. Diabetologia 2015, 58, 1394–1408. [CrossRef]
[PubMed]
53. Du, H.; van der, A.D.L.; Boshuizen, H.C.; Forouhi, N.G.; Wareham, N.J.; Halkjaer, J.; Tjønneland, A.;
Overvad, K.; Jakobsen, M.U.; Boeing, H.; et al. Dietary fiber and subsequent changes in body weight and
waist circumference in European men and women. Am. J. Clin. Nutr. 2010, 91, 329–336. [CrossRef] [PubMed]
54. Sivaprakasam, S.; Prasad, P.D.; Singh, N. Benefits of short-chain fatty acids and their receptors in
inflammation and carcinogenesis. Pharmacol. Ther. 2016, 164, 144–151. [CrossRef] [PubMed]
146
Nutrients 2018, 10, 1590
55. Hjorth, M.F.; Blædel, T.; Bendtsen, L.Q.; Lorenzen, J.K.; Holm, J.B.; Kiilerich, P.; Roager, H.M.; Kristiansen, K.;
Larsen, L.H.; Astrup, A. Prevotella-to-Bacteroides ratio predicts body weight and fat loss success on 24-week
diets varying in macronutrient composition and dietary fiber: Results from a post-hoc analysis. Int. J. Obes.
2018. [CrossRef] [PubMed]
56. Chen, T.T.; Long, W.M.; Zhang, C.H.; Liu, S.; Zhao, L.P.; Hamaker, B.R. Fiber-utilizing capacity varies in
Prevotella- versus Bacteroides-dominated gut microbiota. Sci. Rep. 2017, 7, 2594. [CrossRef] [PubMed]
57. Wu, G.D.; Chen, J.; Hoffmann, C.; Bittinger, K.; Chen, Y.Y.; Keilbaugh, S.A.; Bewtra, M.; Knights, D.;
Walters, W.A.; Knight, R. Linking long-term dietary patterns with gut microbial enterotypes. Science 2011,
334, 105–108. [CrossRef] [PubMed]
58. De Moraes, A.C.F.; Fernandes, G.R.; da Silva, I.T.; Almeida-Pititto, B.; Gomes, E.P.; Pereira, A.D.;
Ferreira, S.R.G. Enterotype May Drive the Dietary-Associated Cardiometabolic Risk Factors. Front. Cell.
Infect. Microbiol. 2017, 7, 47. [CrossRef] [PubMed]
59. Arumugam, M.; Raes, J.; Pelletier, E.; Le Paslier, D.; Yamada, T.; Mende, D.R.; Fernandes, G.R.; Tap, J.;
Bruls, T.; Batto, J.M.; et al. Enterotypes of the human gut microbiome. Nature 2011, 473, 174–180. [CrossRef]
[PubMed]
60. Han, J.-L.; Lin, H.-L. Intestinal microbiota and type 2 diabetes: From mechanism insights to therapeutic
perspective. World J. Gastroenterol. 2014, 20, 17737–17745. [CrossRef] [PubMed]
61. Korpela, K. Diet, Microbiota, and Metabolic Health: Trade-Off Between Saccharolytic and Proteolytic
Fermentation. Annu. Rev. Food Sci. Technol. 2018, 9, 65–84. [CrossRef] [PubMed]
62. Seth, E.C.; Taga, M.E. Nutrient cross-feeding in the microbial world. Front. Microbiol. 2014, 5, 350. [CrossRef]
[PubMed]
63. Loreau, M. Biodiversity and Ecosystem Functioning: Current Knowledge and Future Challenges. Science
2001, 294, 804–808. [CrossRef] [PubMed]
64. Wright, E.M. The Intestinal Na+ /Glucose Cotransporter. Annu. Rev. Physiol. 1993, 55, 575–589. [CrossRef]
[PubMed]
65. Helliwell, P.A.; Richardson, M.; Affleck, J.; Kellett, G.L. Regulation of GLUT5, GLUT2 and intestinal
brush-border fructose absorption by the extracellular signal-regulated kinase, p38 mitogen-activated kinase
and phosphatidylinositol 3-kinase intracellular signalling pathways: Implications for adaptation to diabetes.
Biochem. J. 2000, 350, 163–169. [PubMed]
66. Burant, C.F.; Takeda, J.; Brot-Laroche, E.; Bell, G.I.; Davidson, N.O. Fructose transporter in human
spermatozoa and small intestine is GLUT5. J. Biol. Chem. 1992, 267, 14523–14526. [PubMed]
67. Miyamoto, K. Differential responses of intestinal glucose transporter mRNA transcripts to levels of dietary
sugars. Biochem. J. 1993, 295, 211–215. [CrossRef] [PubMed]
68. Röder, P.V. The role of SGLT1 and GLUT2 in intestinal glucose transport and sensing. PLoS ONE 2014,
9, e89977. [CrossRef] [PubMed]
69. Ferraris, R.P. Dietary and developmental regulation of intestinal sugar transport. Biochem. J. 2001, 360,
265–276. [CrossRef] [PubMed]
70. Gouyon, F. Simple-sugar meals target GLUT2 at enterocyte apical membranes to improve sugar absorption:
A study in GLUT2-null mice. J. Physiol. 2003, 552, 823–832. [CrossRef] [PubMed]
71. Pfannkuche, H.; Gäbel, G. Glucose, epithelium, and enteric nervous system: Dialogue in the dark. J. Anim.
Physiol. Anim. Nutr. 2009, 93, 277–286. [CrossRef] [PubMed]
72. Ferraris, R.P.; Choe, J.-Y.; Patel, C.R. Intestinal Absorption of Fructose. Annu. Rev. Nutr. 2018, 38, 41–67.
[CrossRef] [PubMed]
73. Jones, H.F.; Butler, R.N.; Brooks, D.A. Intestinal fructose transport and malabsorption in humans. Am. J.
Physiol. Gastrointest. Liver Physiol. 2011, 2, 202–206. [CrossRef] [PubMed]
74. Douard, V.; Ferraris, R.P. Regulation of the fructose transporter GLUT5 in health and disease. Am. J. Physiol.
Endocrinol. Metab. 2008, 295, 227–237. [CrossRef] [PubMed]
75. Deal, R.A.; Tang, Y.; Fletcher, R.; Torquati, A.; Omotosho, P. Understanding intestinal glucose transporter
expression in obese compared to non-obese subjects. Surg. Endosc. 2018, 32, 1755–1761. [CrossRef] [PubMed]
76. Pettolino, F.A.; Walsh, C.; Fincher, G.B.; Bacic, A. Determining the polysaccharide composition of plant cell
walls. Nat. Protoc. 2012, 7, 1590–1607. [CrossRef] [PubMed]
77. Valent, B.S.; Albersheim, P. The structure of plant cell walls: On the binding of xyloglucan to cellulose fibers.
Plant. Physiol. 1974, 54, 105–108. [CrossRef] [PubMed]
147
Nutrients 2018, 10, 1590
78. Cummings, J.H. Cellulose and the human gut. Gut 1984, 25, 805–810. [CrossRef] [PubMed]
79. Meibohm, B.; Derendorf, H. Basic concepts of pharmacokinetic/pharmacodynamic (PK/PD) modelling.
Int. J. Clin. Pharmacol. Ther. 1997, 35, 401–413. [PubMed]
80. Ayorinde, J.; Odeniyi, M.; Balogun-Agbaje, O. Formulation and Evaluation of Oral Dissolving Films of
Amlodipine Besylate Using Blends of Starches With Hydroxypropyl Methyl Cellulose. Polym. Med. 2016, 46,
45–51. [CrossRef] [PubMed]
81. Karrout, Y. In vivo efficacy of microbiota-sensitive coatings for colon targeting: A promising tool for IBD
therapy. J. Control. Release 2015, 197, 121–130. [CrossRef] [PubMed]
82. Chourasia, M.K.; Jain, S.K. Pharmaceutical approaches to colon targeted drug delivery systems. J. Pharm.
Pharm. Sci. 2003, 6, 33–66. [PubMed]
83. Alcantara, C.; Zuniga, M. Proteomic and transcriptomic analysis of the response to bile stress of Lactobacillus
casei BL23. Microbiology 2012, 158, 1206–1218. [CrossRef] [PubMed]
84. Moodley, C.; Reid, S.J.; Abratt, V.R. Molecular characterisation of ABC-type multidrug efflux systems in
Bifidobacterium longum. Anaerobe 2015, 32, 63–69. [CrossRef] [PubMed]
85. Valero, T. Mitochondrial biogenesis: Pharmacological approaches. Curr. Pharm. Des. 2014, 20, 5507–5509.
[CrossRef] [PubMed]
86. Martín, R.; Martín, C.; Escobedo, S.; Suárez, J.E.; Quirós, L.M. Surface glycosaminoglycans mediate adherence
between HeLa cells and Lactobacillus salivarius Lv72. BMC Microbiol. 2013, 13. [CrossRef] [PubMed]
87. Neis, E.; Dejong, C.; Rensen, S. The Role of Microbial Amino Acid Metabolism in Host Metabolism. Nutrients
2015, 7, 2930–2946. [CrossRef] [PubMed]
88. Grześkowiak, Ł. The effect of growth media and physical treatments on the adhesion properties of canine
probiotics. J. Appl. Microbiol. 2013, 115, 539–545. [CrossRef] [PubMed]
89. Arakawa, K. Lactobacillus gasseri requires peptides, not proteins or free amino acids, for growth in milk.
J. Dairy Sci. 2015, 98, 1593–1603. [CrossRef] [PubMed]
90. Robitaille, G.; Champagne, C.P. Growth-promoting effects of pepsin- and trypsin-treated caseinomacropeptide
from bovine milk on probiotics. J. Dairy Res. 2014, 81, 319–324. [CrossRef] [PubMed]
91. Ito, T.; Sekizuka, T.; Kishi, N.; Yamashita, A.; Kuroda, M. Conventional culture methods with commercially
available media unveil the presence of novel culturable bacteria. Gut Microbes 2018, 17, 1–15. [CrossRef]
[PubMed]
92. Elhenawy, W.; Debelyy, M.O.; Feldman, M.F. Preferential packing of acidic glycosidases and proteases into
Bacteroides outer membrane vesicles. mBio 2014, 5, e00909-14. [CrossRef] [PubMed]
93. Gandhi, A.; Shah, N.P. Cell growth and proteolytic activity of Lactobacillus acidophilus, Lactobacillus
helveticus, Lactobacillus delbrueckii ssp. bulgaricus, and Streptococcus thermophilus in milk as affected by
supplementation with peptide fractions. Int. J. Food Sci. Nutr. 2014, 65, 937–941. [CrossRef] [PubMed]
94. Nemoto, T.K. Identification of a new subtype of dipeptidyl peptidase 11 and a third group of the S46-family
members specifically present in the genus Bacteroides. Biochimie 2018, 147, 25–35. [CrossRef] [PubMed]
95. Sabljić, I. Crystal structure of dipeptidyl peptidase III from the human gut symbiont Bacteroides
thetaiotaomicron. PLoS ONE 2017, 12, e0187295. [CrossRef] [PubMed]
96. Zhang, Q. Vildagliptin increases butyrate-producing bacteria in the gut of diabetic rats. PLoS ONE 2017,
12, e0184735. [CrossRef] [PubMed]
97. Meyer, H.W. Extensive Resection of Small and Large Intestine: A Further Twenty-Two Year Follow-Up
Report. Ann. Surg. 1962, 168, 287–289. [CrossRef]
98. Garai, P.; Chandra, K.; Chakravortty, D. Bacterial peptide transporters: Messengers of nutrition to virulence.
Virulence 2017, 8, 297–309. [CrossRef] [PubMed]
99. Steiner, H.Y.; Naider, F.; Becker, J.M. The PTR family: A new group of peptide transporters. Mol. Microbiol.
1995, 16, 825–834. [CrossRef] [PubMed]
100. Paulsen, I.T.; Skurray, R.A. The POT family of transport proteins. Trends Biochem. Sci. 1994, 19, 404. [CrossRef]
101. DiRusso, C.C.; Black, P.N. Long-chain fatty acid transport in bacteria and yeast. Paradigms for defining
the mechanism underlying this protein-mediated process. Mol. Cell. Biochem. 1999, 192, 41–52. [CrossRef]
[PubMed]
102. Dirusso, C.C.; Black, P.N. Bacterial long chain fatty acid transport: Gateway to a fatty acid-responsive
signaling system. J. Biol. Chem. 2004, 279, 49563–49566. [CrossRef] [PubMed]
148
Nutrients 2018, 10, 1590
103. Englyst, H.N.; Hay, S.; Macfarlane, G.T. Polysaccharide breakdown by mixed populations of human faecal
bacteria. FEMS Microbiol. Lett. 1987, 45, 163–171. [CrossRef]
104. Round, J.L.; Palm, N.W. Causal effects of the microbiota on immune-mediated diseases. Sci. Immunol. 2018,
20, 1603. [CrossRef] [PubMed]
105. Lu, M.C.; Yan, S.T.; Yin, W.Y.; Koo, M.; Lai, N.-S. Risk of rheumatoid arthritis in patients with type 2 diabetes:
A nationwide population-based case-control study. PLoS ONE 2014, 7, e101528. [CrossRef] [PubMed]
106. Ehrlich, S.F.; Quesenberry, C.P.; Van Den Eeden, S.K.; Shan, J.; Ferrara, A. Patients Diagnosed With
Diabetes Are at Increased Risk for Asthma, Chronic Obstructive Pulmonary Disease, Pulmonary Fibrosis,
and Pneumonia but Not Lung Cancer. Diabetes Care 2010, 33, 55–60. [CrossRef] [PubMed]
107. Lønnberg, A.S. Association of Psoriasis With the Risk for Type 2 Diabetes Mellitus and Obesity.
JAMA Dermatology 2016, 57, 645–659. [CrossRef] [PubMed]
108. Wojciechowska, J.; Krajewski, W.; Bolanowski, M.; Kr˛ecicki, T.; Zatoński, T. Diabetes and Cancer: A Review
of Current Knowledge. Exp. Clin. Endocrinol. Diabetes 2016, 124, 263–275. [CrossRef] [PubMed]
109. Holden, S.E. Diabetes and Cancer. Endocr. Dev. 2016, 31, 135–145. [PubMed]
110. Faurschou, A. Improvement in psoriasis after treatment with the glucagon-like peptide-1 receptor agonist
liraglutide. Acta Diabetol. 2014, 51, 147–150. [CrossRef] [PubMed]
111. Al-Badri, M.R.; Azar, S.T. Effect of glucagon-like peptide-1 receptor agonists in patients with psoriasis.
Ther. Adv. Endocrinol. Metab. 2014, 5, 34–38. [CrossRef] [PubMed]
112. Ahern, T. Glucagon-like peptide-1 analogue therapy for psoriasis patients with obesity and type 2 diabetes:
A prospective cohort study. J. Eur. Acad. Dermatol. Venereol. 2013, 27, 1440–1443. [CrossRef] [PubMed]
113. Schwartz, A.V. Diabetes, bone and glucose-lowering agents: Clinical outcomes. Diabetologia 2017, 7,
1170–1179. [CrossRef] [PubMed]
114. Malik, V.S.; Popkin, B.M.; Bray, G.A.; Després, J.P.; Hu, F.B. Sugar-Sweetened Beverages, Obesity, Type 2
Diabetes Mellitus, and Cardiovascular Disease Risk. Circulation 2010, 121, 1356–1364. [CrossRef] [PubMed]
115. Kim, S.C.; Schneeweiss, S.; Glynn, R.J.; Doherty, M.; Goldfine, A.B.; Solomon, D.H. Dipeptidyl peptidase-4
inhibitors in type 2 diabetes may reduce the risk of autoimmune diseases: A population-based cohort study.
Ann. Rheum. Dis. 2015, 74, 1968–1975. [CrossRef] [PubMed]
116. Shi, Z. Association between soft drink consumption and asthma and chronic obstructive pulmonary disease
among adults in Australia. Respirology 2012, 17, 363–369. [CrossRef] [PubMed]
117. DeChristopher, L.R.; Uribarri, J.; Tucker, K.L. Intakes of apple juice, fruit drinks and soda are associated with
prevalent asthma in US children aged 2-9 years. Public Health Nutr. 2016, 19, 123–130. [CrossRef] [PubMed]
118. Legaki, E.; Gazouli, M. Influence of environmental factors in the development of inflammatory bowel
diseases. World J. Gastrointest. Pharmacol. Ther. 2016, 7, 112–125. [CrossRef] [PubMed]
119. Bomback, A.S.; Derebail, V.K.; Shoham, D.A.; Anderson, C.A.; Steffen, L.M.; Rosamond, W.D.K.A.
Sugar-sweetened soda consumption, hyperuricemia, and kidney disease. Kidney Int. 2012, 77, 609–616.
[CrossRef] [PubMed]
120. De Koning, L. Sweetened beverage consumption, incident coronary heart disease, and biomarkers of risk in
men. Circulation 2012, 125, 1735–1741. [CrossRef] [PubMed]
121. Bunn, H.F.; Higgins, P.J. Reaction of monosaccharides with proteins: Possible evolutionary significance.
Science 1981, 213, 222–224. [CrossRef] [PubMed]
122. Ahmed, N. Advanced glycation endproducts—Role in pathology of diabetic complications. Diabetes Res.
Clin. Pract. 2005, 67, 3–21. [CrossRef] [PubMed]
123. Ramasamy, R.; Yan, S.F.; Schmidt, A.M. The diverse ligand repertoire of the receptor for advanced glycation
endproducts and pathways to the complications of diabetes. Vascul. Pharmacol. 2012, 57, 160–167. [CrossRef]
[PubMed]
124. Chen, Y.-S.; Yan, W.; Geczy, C.L.; Brown, M.A.; Thomas, R. Serum levels of soluble receptor for advanced
glycation end products and of S100 proteins are associated with inflammatory, autoantibody, and classical
risk markers of joint and vascular damage in rheumatoid arthritis. Arthritis Res. Ther. 2009, 11, 39. [CrossRef]
[PubMed]
125. Rao, R.; Sen, S.; Han, B.; Ramadoss, S.; Chaudhuri, G. Gestational diabetes, preeclampsia and cytokine
release: Similarities and differences in endothelial cell function. Adv. Exp. Med. Biol. 2014, 814, 69–75.
[PubMed]
149
Nutrients 2018, 10, 1590
126. Dornadula, S.; Elango, B.; Balashanmugam, P.; Palanisamy, R.; Kunka Mohanram, R. Pathophysiological
insights of methylglyoxal induced type-2 diabetes. Chem. Res. Toxicol. 2015, 28, 1666–1674. [CrossRef]
[PubMed]
127. Wu, L.; Juurlink, B.H.J. Increased Methylglyoxal and Oxidative Stress in Hypertensive Rat Vascular Smooth
Muscle Cells. Hypertension 2002, 39, 809–814. [CrossRef] [PubMed]
128. Cardona, F.; Andrés-Lacueva, C.; Tulipani, S.; Tinahones, F.J.; Queipo-Ortuño, M.I. Benefits of polyphenols
on gut microbiota and implications in human health. J. Nutr. Biochem. 2013, 24, 1415–1422. [CrossRef]
[PubMed]
129. Cahenzli, J.; Köller, Y.; Wyss, M.; Geuking, M.B.; McCoy, K.D. Intestinal microbial diversity during early-life
colonization shapes long-term IgE levels. Cell. Host Microbe 2013, 14, 559–570. [CrossRef] [PubMed]
130. Palm, N.W.; de Zoete, M.R.; Flavell, R.A. Immune-microbiota interactions in health and disease.
Clin. Immunol. 2015, 159, 122–127. [CrossRef] [PubMed]
131. Williams, A.; Flavell, R.A.; Eisenbarth, S.C. The role of NOD-like Receptors in shaping adaptive immunity.
Curr. Opin. Immunol. 2010, 22, 34–40. [CrossRef] [PubMed]
132. Abreu, M.T. Toll-like receptor signalling in the intestinal epithelium: How bacterial recognition shapes
intestinal function. Nat. Rev. Immunol. 2010, 10, 131–144. [CrossRef] [PubMed]
133. Caesar, R. Gut-derived lipopolysaccharide augments adipose macrophage accumulation but is not essential
for impaired glucose or insulin tolerance in mice. Gut 2012, 61, 1701–1707. [CrossRef] [PubMed]
134. Fouts, D.E.; Torralba, M.; Nelson, K.E.; Brenner, D.A.; Schnabl, B. Bacterial translocation and changes in the
intestinal microbiome in mouse models of liver disease. J. Hepatol. 2012, 56, 1283–1292. [CrossRef] [PubMed]
135. Tan, J. The Role of Short-Chain Fatty Acids in Health and Disease. Adv. Immunol. 2014, 121, 91–119.
[PubMed]
136. Corrêa-Oliveira, R.; Fachi, J.L.; Vieira, A.; Sato, F.T.; Vinolo, M.A.R. Regulation of immune cell function by
short-chain fatty acids. Clin. Transl. Immunol. 2016, 5, e73. [CrossRef] [PubMed]
137. Remely, M. Effects of short chain fatty acid producing bacteria on epigenetic regulation of FFAR3 in type 2
diabetes and obesity. Gene 2014, 537, 85–92. [CrossRef] [PubMed]
138. Jandhyala, S.M. Role of the normal gut microbiota. World J. Gastroenterol. 2015, 21, 8787–8803. [CrossRef]
[PubMed]
139. Duca, F.A.; Swartz, T.D.; Sakar, Y.; Covasa, M. Increased oral detection, but decreased intestinal signaling for
fats in mice lacking gut microbiota. PLoS ONE 2012, 7, e39748. [CrossRef] [PubMed]
140. Waldecker, M.; Kautenburger, T.; Daumann, H.; Busch, C.; Schrenk, D. Inhibition of histone-deacetylase
activity by short-chain fatty acids and some polyphenol metabolites formed in the colon. J. Nutr. Biochem.
2008, 19, 587–593. [CrossRef] [PubMed]
141. Schilderink, R. The SCFA butyrate stimulates the epithelial production of retinoic acid via inhibition of
epithelial HDAC. Am. J. Physiol. Liver Physiol. 2016, 310, G1138–G1146. [CrossRef] [PubMed]
142. Fagarasan, S.; Honjo, T. Intestinal IgA synthesis: Regulation of front-line body defences. Nat. Rev. Immunol.
2003, 3, 63–72. [CrossRef] [PubMed]
143. Schey, R.; Danzer, C.; Mattner, J. Perturbations of mucosal homeostasis through interactions of intestinal
microbes with myeloid cells. Immunobiology 2015, 220, 227–235. [CrossRef] [PubMed]
144. Wlodarska, M.; Willing, B.P.; Bravo, D.M.; Finlay, B.B. Phytonutrient diet supplementation promotes
beneficial Clostridia species and intestinal mucus secretion resulting in protection against enteric infection.
Sci. Rep. 2015, 5, 9253. [CrossRef] [PubMed]
145. Luo, J. Phytonutrient genistein is a survival factor for pancreatic β-cells via GPR30-mediated mechanism.
J. Nutr. Biochem. 2018, 58, 59–70. [CrossRef] [PubMed]
146. Fu, Z. Genistein Induces Pancreatic β-Cell Proliferation through Activation of Multiple Signaling Pathways
and Prevents Insulin-Deficient Diabetes in Mice. Endocrinology 2010, 151, 3026–3037. [CrossRef] [PubMed]
147. WHO (World Health Organization). Mortality and Burden of Disease Attributable to Selected Major Risks; WHO:
Geneva, Switzerland, 2009.
© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
150
nutrients
Review
Glutamine: Metabolism and Immune Function,
Supplementation and Clinical Translation
Vinicius Cruzat 1,2, *, Marcelo Macedo Rogero 3 , Kevin Noel Keane 1 , Rui Curi 4 and
Philip Newsholme 1, *
1 School of Pharmacy and Biomedical Sciences, Curtin Health Innovation Research Institute, Biosciences,
Curtin University, Perth 6102, Australia; Kevin.Keane@curtin.edu.au
2 Faculty of Health, Torrens University, Melbourne 3065, Australia
3 Department of Nutrition, Faculty of Public Health, University of São Paulo, Avenida Doutor Arnaldo 715,
São Paulo 01246-904, Brazil; mmrogero@usp.br
4 Interdisciplinary Post-Graduate Program in Health Sciences, Cruzeiro do Sul University,
São Paulo 01506-000, Brazil; ruicuri59@gmail.com
* Correspondence: Vinicius.cruzat@laureate.edu.au (V.C.); Philip.newsholme@curtin.edu.au (P.N.)
Abstract: Glutamine is the most abundant and versatile amino acid in the body. In health and
disease, the rate of glutamine consumption by immune cells is similar or greater than glucose.
For instance, in vitro and in vivo studies have determined that glutamine is an essential nutrient
for lymphocyte proliferation and cytokine production, macrophage phagocytic plus secretory
activities, and neutrophil bacterial killing. Glutamine release to the circulation and availability
is mainly controlled by key metabolic organs, such as the gut, liver, and skeletal muscles. During
catabolic/hypercatabolic situations glutamine can become essential for metabolic function, but its
availability may be compromised due to the impairment of homeostasis in the inter-tissue metabolism
of amino acids. For this reason, glutamine is currently part of clinical nutrition supplementation
protocols and/or recommended for immune suppressed individuals. However, in a wide range of
catabolic/hypercatabolic situations (e.g., ill/critically ill, post-trauma, sepsis, exhausted athletes),
it is currently difficult to determine whether glutamine supplementation (oral/enteral or parenteral)
should be recommended based on the amino acid plasma/bloodstream concentration (also known
as glutaminemia). Although the beneficial immune-based effects of glutamine supplementation are
already established, many questions and evidence for positive in vivo outcomes still remain to be
presented. Therefore, this paper provides an integrated review of how glutamine metabolism in key
organs is important to cells of the immune system. We also discuss glutamine metabolism and action,
and important issues related to the effects of glutamine supplementation in catabolic situations.
1. Introduction
At the most basic level, amino acids are the building blocks of proteins in our cells and tissues,
and after water are the second most abundant compound in mammals. Amino acids can be obtained
from endogenous and/or exogenous (i.e., diet) proteins, and their availability is of fundamental
importance for cell survival, maintenance, and proliferation. Mammals, in particular, have developed
biochemical and metabolic pathways to control pathogen infection by increasing amino acid catabolism
to aid immune responses, thus restricting the availability of nitrogen-containing nutrients to invading
microorganisms [1]. This evolutionary mechanism also becomes advantageous for the host to control
its own inflammatory responses to infections.
Among the 20 amino acids detailed in the genetic code, glutamine provides the best example
of the versatility of amino acid metabolism and immune function. Glutamine is the most abundant
and versatile amino acid in the body, and is of fundamental importance to intermediary metabolism,
interorgan nitrogen exchange via ammonia (NH3 ) transport between tissues, and pH homeostasis.
In almost every cell, glutamine can be used as a substrate for nucleotide synthesis (purines, pyrimidines,
and amino sugars), nicotinamide adenine dinucleotide phosphate (NADPH), antioxidants, and many
other biosynthetic pathways involved in the maintenance of cellular integrity and function [2–4].
Most cells in the body function with a constant turnover and/or supply of nutrients, however, cells
of the immune system frequently have to function under nutrient restricted microenvironments [1].
Although glucose is a vital metabolite, and the main fuel for a large number of cells in the body,
cells of the immune system, such as lymphocytes, neutrophils, and macrophages, utilize glutamine
at high rates similar to or greater than glucose under catabolic conditions, such as sepsis, recovery
from burns or surgery, and malnutrition, as well as high intensity/volume physical exercise [5,6].
This theory was first experimentally confirmed in the 1980’s by the laboratory of Eric Newsholme
(1935–2011, widely accepted now as the origin of hypotheses and evidence for the concept of
“immunometabolism”) [3,7] in the University of Oxford, and subsequently by many other laboratories
worldwide [4,8,9]. For this reason, glutamine is considered as a “fuel for the immune system”, where a
low blood concentration may impair immune cell function, resulting in poor clinical outcomes and
increased risk of mortality [10].
Currently, glutamine is routinely supplied as a component of clinical nutrition supplementation
for pre-and post-operative patients, and also for many elite athletes to restore immune functions.
Although there is a growing evidence in support of the immune mediating effects of glutamine
supplementation, several questions and specific considerations still remain. Therefore, the aim of the
present study is to provide an integrated review on how glutamine metabolism in key organs, such as
the gut, liver, and skeletal muscles, is important to cells of the immune system. These key organs
control glutamine availability and shed light on important considerations in regards to glutaminemia
(glutamine concentration into the bloodstream). The immune-enhancing properties and related
paradigms of glutamine supplementation in health and disease are also discussed herein.
152
Nutrients 2018, 10, 1564
primarily glutamine-consuming tissues, such as the intestinal mucosa, leukocytes, and renal tubule
cells, have high glutaminase activity and cofactors capable of degrading glutamine. However, the liver
may become a glutamine-consuming site, and tissues, such as muscle tissue, may present reduced
glutamine synthesis under certain conditions, such as reduced carbohydrate [16] and/or amino
acid [17] intake, high catabolic situations, and/or diseases and stress [5]. Many other factors—mainly
glucocorticoids [18], thyroid hormones [19], growth hormone [20], and insulin [21]—can modulate the
activity performed by glutamine metabolism-regulating enzymes.
Several enzymes are involved in glutamine metabolism; the two main intracellular enzymes
are glutamine synthetase (GS, EC 6.3.1.2) and phosphate-dependent glutaminase (GLS, EC 3.5.1.2).
GS is responsible for triggering the reaction that synthesizes glutamine from ammonium ion (NH4+ ),
and glutamate through ATP consumption, whereas GLS is responsible for glutamine hydrolysis,
which converts it into glutamate and NH4+ again [22,23] (Figure 1). With respect to the intracellular
location, GS is primarily found in the cytosol, whereas GLS (in its active form) is mainly found within
the mitochondria. These locations are compatible with the enzymes’ functions: GS produces glutamine
for the synthesis of cytoplasmic proteins and nucleotides, whereas GLS catalyses glutamine conversion
to glutamate as an important step to the tricarboxylic acid cycle (TCA, also known as the Krebs cycle)
entry at 2-oxoglutarate as an energy source or source of metabolic intermediates [3].
Figure 1. Glutamine synthesis and hydrolysis. Glutamine is mainly synthesized by the enzyme
glutamine synthetase (GS) and hydrolysed by the enzyme, glutaminase (GLS). GS catalyses glutamine
biosynthesis using glutamate and ammonia (NH3 ) as a source. In this reaction, one ATP is consumed.
Glutamate can be donated by many amino acids obtained exogenously (i.e., diet) and/or from
endogenous amino acids’ catabolism. On the other hand, GLS is responsible for glutamine hydrolysis
to glutamate and ammonium ion (NH4 ). Almost all cells in the body express GS and GLS, and their
predominant expression and activity will dictate if the tissue is more likely to produce or consume
glutamine in health and disease.
153
Nutrients 2018, 10, 1564
Glutamine tissue and blood concentrations are dependent on GS or GLS activities. Endogenous
glutamine synthesis does not meet the human body’s demands in catabolic conditions, such as in
cancer [25], sepsis [4,26], infections [27,28], surgeries [8], traumas [10], as well as during intense and
prolonged physical exercise [29,30]. Glutamine assumes the role of a conditionally essential amino acid
in such deficiency conditions by promoting a concomitant increase in GLS expression and inhibiting
the GS action [14]. However, it is worth emphasizing that, although plasma glutamine concentration
is reduced from its normal concentration (i.e., 500–800 μmol/L) to 300–400 μmol/L, cells depending
on this amino acid, such as immune cells, are in fact poorly influenced in terms of proliferation
and function [6]. On the other hand, the high tissue catabolism leads to reduced glutamine stock
in human tissues, mainly in the muscle and liver (Figure 3). The low glutamine concentration in
human tissues affects the whole body since this amino acid provides nitrogen atoms to the synthesis
of purines, pyrimidines, and amino sugars [31]. If the high glutamine degradation in these tissues
persists, a large number of metabolic pathways and mechanisms that depend on glutamine availability
are affected, resulting in immunosuppression. More recently, studies reported that bacterial infections
154
Nutrients 2018, 10, 1564
(e.g., Escherichia coli) can alter its metabolism and harness glutamine to suppress the effects of acid
stress and copper toxicity [32]. Hence, bacterial pathogens can adapt and survive by altering core
metabolic pathways important for host-imposed antibacterial strategies.
Figure 3. Glutamine inter-tissue metabolic flux starting in skeletal muscle, liver, and gut continues
in the immune cells. Abbreviations: Glutamine, GLN; glutamate, GLU; aspartate, ASP; arginine,
ARG; leucine, LEU; alanine, ALA; glucose, Gluc; pyruvate, Pyr; pyruvate dehydrogenase; PDC;
pyruvate carboxylase, PC; malate dehydrogenase, MD; glyceraldehyde-3-Phosphate, G3-P; lactate, Lac;
triacylglycerol, TG; ribose 5-phosphate, R5P; alanine aminotransferase, ALT; glutamate dehydrogenase,
GDH; glutamine synthetase, GS; glutaminase, GLS; inducible nitric oxide synthase, iNOS; intracellular
heat shock protein, iHSP; heat Shock Factor 1, HSF-1; heat shock elements, HSEs; sirtuin 1, SIRT1;
hexosamine biosynthetic pathway, HBP; ammonia, NH3 ; glutathione, GSH; oxidized GSH, GSSG;
glutathione S-reductase, GSR; protein kinase B, Akt; AMP-activated protein kinase, AMPK; mTOR
complex 1 and 2, mTORC1/2, extracellular signal-regulated kinases, ERK; c-Jun N-terminal kinases,
JNK; gamma-Aminobutyric acid, GABA.
155
Nutrients 2018, 10, 1564
The glutamine hydrolysis into glutamate, which is catalysed by GLS, corresponds to the first
reaction resulting from glutamine consumption [36]. Although the gut is the major site of glutamine
consumption, glutamine concentration in the intestinal tissue is low. This is due to the high GLS
activity (3–6 μmol/hour/mg of protein), and also high GLS affinity for the substrate, glutamine.
Interestingly, there is a correlation between the presence of GSL and the use of glutamine by certain
cell types. Almost all GLS found in the intestinal cells is bound to the mitochondrial membrane.
The modulation of GLS activity in the intestinal tissue is important to maintain the tissue integrity
and enable adequate absorption of nutrients, as well as to prevent bacterial translocation into the
bloodstream (i.e., septicaemia) [37]. Prolonged fasting and malnutrition states are associated with
reduced GLS activity; on the other hand, GLS activity increases in the postprandial period, after the
administration of enteral feeding of branched-chain amino acids and/or L-alanyl-L-glutamine [38].
The ATP-ubiquitin-dependent proteolytic pathway associated with proteasome is known to
degrade endogenous short-lived or abnormal proteins/peptides, as well as participating in the
regulation of the inflammatory response. The ATP-ubiquitin-dependent proteolytic pathway could be
important for the turnover of gut mucosal proteins, which are very short lived. Indeed, the nuclear
factor of kappa light polypeptide gene enhancer in B-cells inhibitor (IκB) ubiquitinylation allows the
nuclear factor kappa-light-chain-enhancer of activated B cells’ (NF-κB) translocation in the nucleus,
and thus transcription of proinflammatory genes [17,39,40] (Figure 3). Glutamine stimulates protein
synthesis and reduces ubiquitin-dependent proteolysis in the enterocyte since this amino acid reduces
ubiquitin gene expression. Glutamine can increase the gene expression of the arginine-succinate
synthase enzyme in Caco-2 cells (human colon epithelial-line cell). Glutamine activates the extracellular
signal-regulated kinases (ERKs) and the c-Jun N-terminal kinases (JNK) in the enterocyte, and it leads
to significant increase in c-Jun gene expression and in the activity of the transcription factor known
as activator protein 1 (AP-1) [41,42]. Such glutamine action potentiates the effects of growth factors
on cell proliferation and repair. Heat shock (43 ◦ C) induces intestinal epithelial cell death, which can
be exacerbated due to the lack of glutamine. However, as it happens with muscle tissues, glutamine
supplementation enables a dose-dependent reduction in heat shock-associated cell death. This effect of
glutamine partly results from the amino acid capacity of increasing the gene expression of HSP 70 [37].
Dysregulation of cytokine production plays a major role in the pathogenesis of inflammatory
bowel disease (IBD). The gut mucosa of patients with IBD (Crohn’s disease or ulcerative colitis) has
been reported to produce high amounts of proinflammatory cytokines, such as interleukin (IL-)1β,
IL-6, IL-8, and tumour necrosis factor-alpha (TNF-α), in contrast to a less marked increase in the
production of anti-inflammatory cytokines, such as IL-10. For example, Coeffier, et al. [43] verified that
glutamine reduces pro-inflammatory cytokine production by human intestinal mucosa, probably by a
posttranscriptional pathway. Glutamine could be useful to modulate inflammatory conditions with
imbalanced cytokine production.
156
Nutrients 2018, 10, 1564
Hormones, such as insulin and insulin-like growth factors (IGFs), stimulate glutamine transport
into the intracellular environment, whereas glucocorticoids stimulate glutamine release into the
extracellular space. The transmembrane gradient for glutamine through the skeletal muscle is high,
a fact that restricts amino acid diffusion through the cell membrane. Glutamine is actively transported
into cells through a sodium-dependent channel system, whose outcome is a consumption of ATP.
The glutamine transport through the muscle cell membrane is faster than the transport of all other
amino acids [48]. Interestingly, the constant maintenance of glutamine availability in the intracellular
fluid, as well as the high glutamine concentration gradient across the cell membrane, is supported by
many pathways, such as the transport system affinity for the amino acid, its intracellular turnover
ratio and the extracellular supply, the intra- and extracellular sodium concentrations, and the influence
of other amino acids competing for carrier molecules [49,50].
During the post-absorptive state, approximately 50% of the glutamine synthesis in the skeletal
muscle takes place through glutamate uptake from the bloodstream, a fact that characterizes part of
the glutamine-glutamate cycle. In addition, muscle protein catabolism directly produces glutamine,
although it also leads to BCAAs, glutamate, aspartate, and asparagine release. The carbon skeletons of
these amino acids are used for glutamine de novo synthesis [51,52]. Glutamine and alanine correspond,
respectively, to 48% and 32% of the amino acids released by the skeletal muscle in the post-absorptive
state; the glutamine containing two nitrogen atoms per molecule is the main muscle nitrogen-release
source. The glutamine and alanine exchange rates exceed their abundance in the body, and their
occurrence in the muscle protein corresponds to 10–15%, thus indicating the constant need of glutamine
and alanine de novo synthesis in the skeletal muscle [4]. The glutamine synthesis rate in the skeletal
muscle (approximately 50 mmol/h) is higher than that recorded for any other amino acid. Thus,
glutamine and alanine should result from the interconversion of amino acids within the cell, in a
process that depends on the cell metabolic conditions, which are affected by human nutritional and
hormonal status, as well as by physical exercise [53,54].
One of the first studies about muscle glutamine metabolism in catabolic situations has recorded
that reduced glutamine concentration in the skeletal muscle is associated with the reduced survival
rate of sepsis-state patients. The severe muscle glutamine-concentration decrease in critically-ill
patients (80% reduction, on average, in the normal concentration due to protein degradation) is
accompanied by increased glutamine synthesis and release from the skeletal muscle. It happens
because of the increased messenger RNA (mRNA) and GS enzyme activity in the skeletal muscle
during severe catabolic states. Glucocorticoids may increase the amount of GS mRNA in the muscle
tissue through a glucocorticoid receptor-dependent process that happens in the cytosol. Once the
glucocorticoid is bound to its cytosolic receptor they are translocated to the nucleus, where they
bind to regions containing glucocorticoid-response elements, which induce GS gene transcription,
among others [55,56].
Although the GS activity increases in response to physiological stress, the protein amount may
not increase in parallel to that of the mRNA, thus indicating the activation of post-transcriptional
control mechanisms. Thus, the activity of the aforementioned enzyme appears to be controlled through
the intracellular glutamine concentration by means of a post-transcriptional control mechanism,
which increases the GS enzyme activity when the intracellular glutamine concentration decreases.
However, the GS enzyme is relatively unstable in the presence of glutamine; therefore, the increased
intracellular glutamine concentration leads to faster GS degradation. In addition, glucocorticoids and
intracellular glutamine depletion work synergistically by increasing the GS expression in the skeletal
muscle [57].
In vitro studies conducted with several cell types demonstrated that glutamine can also change the
gene expression of contractile proteins. According to one study, cardiomyocyte growth and maturation
were accompanied by increased mRNA contents in proteins, such as alpha-myosin heavy-chain
(α-MHC) and alpha-actin; both parameters were considered non-pathologic hypertrophy [58].
Other studies highlight the relevant role played by glutamine in mediating the activation of pathways,
157
Nutrients 2018, 10, 1564
such as the mammalian target of rapamycin (mTOR), which is considered an essential tissue size and
mass regulator, either in healthy or ill patients. In fact, the use of amino acids, mainly of leucine,
as anabolic inducers in muscle cells has its action compromised via mTOR when glutamine is not
available [17]. Despite the essential role played by glutamine in regulating the expression of muscle
content-associated genes, there are no in vivo studies supporting the hypothesis that supplementation
applied alone can promote muscle mass increase.
Another significant role played by glutamine is associated with its capacity of modulating
protective and resistance responses to injuries, which are also known as antioxidant and cytoprotective
effects. The high oxidative stress generated in catabolism situations results in several effects that
culminate in pro-apoptotic stimuli through classic pathways, such as that of the NF-κB. Reactive oxygen
species (ROS), both the radical and the non-radical species, react to minerals, to phospholipid
membranes, and to proteins, among other relevant compounds, to cellular homeostasis [59].
Glutamine can modulate the expression of heat shock proteins (HSP). According to a study conducted
with acutely-inflamed mice (subjected to endotoxemia, which is a sepsis model), the increased
glutamine availability in the animals’ tissues helped to keep the HSP expression, mainly in the
70 (the most abundant form), 90, and 27 kDa family. Results concerning skeletal and liver muscles were
recorded at protein and gene expression levels. In addition, other genes were highly responsive to
glutamine, such as the heat shock factor 1 (HSF-1), which is important for HSP synthesis, and enzymes
linked to the antioxidant system (Figure 3). The glutamate resulting from glutamine is an essential
substrate for glutathione synthesis, a fact that changes the expression of genes, such as glutathione
S-reductase (GSR) and glutathione peroxidase 1 (GPx1). The glutamine cytoprotective and antioxidant
properties may be particularly important in high catabolism situations, in which the activity and the
expression of inflammatory pathways mediated by NF-κB are modulated [4,39].
158
Nutrients 2018, 10, 1564
regulate flux through CP generation. High glutamine levels increase glutamate and NH3 through GLS
activity, with the NH3 reacting with ornithine to form CP. NH3 is also a co-activator of GLS in the
liver [62,63], and since liver GLS is not inhibited by glutamate [61,64], the glutamate produced can
form N-acetylglutamate, which is an activator of both GLS and CPS [65,66]. Consequently, both of
these biochemical mechanisms result in a high flux toward urea formation due to the elevated levels of
NH3 in the mitochondrial (sourced from digestive tract blood and glutamine degradation), the high
affinity of CPS for NH3 , and the increased activity of GLS. However, the ultimate generation of urea
appears to be regulated by the sub-cellular level of glutamine and, consequently, its uptake from the
extracellular environment.
The intra-, inter-, and extracellular transport of glutamine in hepatocytes is central to the
conversion of the NH3 /NH4 + to urea, its subsequent excretion, and blood pH balance due to effects
on HCO3 − . The liver architecture is exquisitely designed such that periportal hepatocytes (near the
portal vein) receiving blood and nutrients from the gut are primarily responsible for urea production
using glutamine as outlined above. However, the distal hepatocytes near the hepatic vein (perivenous),
utilise any remaining NH4 + from the blood, including that bypassing the periportal hepatocytes, for
the re-synthesis and secretion of glutamine into the circulation. In this latter scenario, the NH3 /NH4 +
enters the perivenous hepatocytes and, using glutamate as a substrate, glutamine is generated by
GS. This process scavenges any NH3 /NH4 + escaping the periportal process, while also replenishing
the glutamine that was used by periportal hepatocytes in the generation of urea and disposal of
nitrogen. The different functional regions of the liver were illustrated by the expression status of
glutamine-metabolizing enzymes, such that high levels of GLS were found in the portal region [67,68],
while only 7% of hepatocytes expressed GS, and these were found specifically around the central
hepatic veins [69].
The intercellular or liver compartmentalized cycling of glutamine between these liver regions
is also mediated by specific membrane transporters in periportal and perivenous hepatocytes.
These transport systems also control intra- and extracellular pH by the antiport translocation of
H+ on the entry of Na+ and glutamine. Glutamine from the diet is taken up by periportal hepatocytes
along with two Na+ ions, while one H+ is extruded in the opposite direction from the hepatocyte into
the extracellular space. This process is driven by the concentrations of glutamine and Na+ outside
the cell (i.e., in the blood), and also the intracellular concentration of H+ . In essence, this directional
transport is also regulated by the relative pH difference between the intracellular and extracellular
space, such that periportal hepatic glutamine uptake leads to extracellular acidification and intracellular
alkalization, while glutamine export from perivenous hepatocytes leads to intracellular acidification
and extracellular alkalization [61]. Experiments in perfused rat livers have shown that a slight alkali
increase in extracellular pH (0.4 units) can enhance mitochondrial import of glutamine into hepatocytes,
as H+ is transported externally to possibly reduce extracellular pH and maintain equilibrium.
Mitochondrial glutamine concentration increased to about 15–50 mM, while the extracellular (0.6 mM)
and cytosolic (6 mM) glutamine concentrations remain unchanged in these perfusion models [70,71].
Therefore, the regulation and flux through GLS in periportal hepatocytes are regulated by the
sub-cellular concentration of glutamine, and not just the rate of glutamine entry from extracellular
sources. For perivenous hepatocytes, the synthesis and release of glutamine to the blood is facilitated
by the increased cytoplasmic, and lower extracellular concentration of glutamine, along with the less
acidic intracellular environment, which is countered by the antiport cytosolic influx of H+ [61].
Importantly, glutamine importation and exportation also affect osmotic balance and therefore
influences hepatocyte volume. This has additional consequences for hepatic function, including bile
synthesis and release [72], but also regulates anabolic processes, such as glycogen, lipid, and protein
synthesis [61]. Largely, glutamine uptake enhances hepatocyte cell swelling and hydration [73],
which leads to increased glycogen and fatty acid synthesis [74,75], and reduced proteolysis mediated
by P38 mitogen-activated protein kinases (p38 MAPK) signalling [76]. Other amino acids, such as
glycine and alanine, along with anabolic hormones, such as insulin, promoted hepatocyte swelling,
159
Nutrients 2018, 10, 1564
leading to increased biosynthetic processes [77], while catabolic hormones, like glucagon, reduced
intracellular glutamine levels induced hepatocyte shrinkage [78]. Consequently, dehydration due
to reduced intracellular glutamine levels is characterized by decreased cell volume, initiation of the
catabolic process, and insulin-resistant conditions, and it was recently shown that hypertonic infusion
can cause glucose dysregulation in humans [79].
The liver is an insulin-sensitive organ and like skeletal muscle, is responsible for glucose disposal
via glycogen synthesis. Development of insulin resistance and subsequent glucotoxic conditions can
progress to chronic disorders, such as non-alcoholic fatty liver disease (NAFLD), characterized by
excessive lipid accumulation, and non-alcoholic steatohepatitis (NASH), characterized by increased
extracellular matrix (ECM) deposition [80,81]. These chronic disorders may lead to further hepatocyte
damage, manifesting as liver cirrhosis and possibly hepatocellular carcinoma. The liver can be
damaged in various ways, including infection (e.g., hepatitis B and C), alcoholism, metabolic disease,
and prolonged unhealthy diets. This damage elicits a pro-inflammatory hepatic environment,
which leads to liver tissue fibrosis, causing impaired hepatic function [80]. Untreated fibrosis ultimately
progresses to cirrhosis, which is mostly irreversible [80]. A key mediator of liver fibrosis is the hepatic
stellate cell (HSC), which is a mesenchymal, fibrogenic cell that resides in the sub-endothelial space
of Disse between the hepatocyte epithelium and the sinusoids. While normally in a quiescent state,
these cells become activated following liver insult, and they respond to cytokines and proliferate
to aid injury repair. However, overactivation or a failure to resolve their activation status (chronic
activation from continued exposure to pro-inflammatory stimuli) can lead to the increased ECM
deposition in the space of Disse that has a negative consequence for hepatocyte function and normal
liver architecture, such as loss of microvilli [80,82]. Kupffer cells, a liver macrophage, are also activated
in these conditions, and together with HSCs promote a pro-inflammatory hepatic environment.
Pro-inflammatory activators of these cells are beyond the scope of this manuscript, but it has been
demonstrated that HSCs require glutamine metabolism to maintain proliferation. It was shown
that activated HSC were dependent on glutamine conversion to ∝-ketoglutarate and non-essential
amino acids for proliferation, and reduction of glutamine significantly impaired HSC activation [82].
In addition, glutamine can be used as a precursor for proline synthesis, which is a key component of
collagen and ECM formation [82].
At present, there is some evidence to indicate that glutamine supplements slow NAFLD [83] or
NASH [84] progression, but most studies have been conducted in rodents. There is no convincing
evidence to indicate that glutamine supplements prevent NAFLD or NASH progression in humans,
which may be due to the complexity of multiple factors contributing to these disorders. The liver is a
remarkable organ and has the ability to regenerate, and some research has indicated that glutamine
supplementation may be advantageous for liver growth and repair after resection, but are again
limited to animal studies [85]. However, others have suggested that raised glutamine levels are
associated with liver failure, and the severity correlated with plasma glutamine [86]. Consequently,
the effects of glutamine on liver function beyond urea synthesis have not been fully explored, and the
administration of exogenous glutamine to those with compromised hepatic function needs to be
considered carefully [86].
160
Nutrients 2018, 10, 1564
measure the levels in plasma and tissues. Throughout the 1960s, 1970s, and 1980s, Hans Krebs, Philip
Randle (1926–2006), Derek Williamson (1929–1998), and Eric Newsholme (1935–2011) all worked on
metabolic regulation utilizing different research models, from isolated cells in vitro, to human and
in vivo experiments. Although glucose is a vital metabolite, and the main fuel for a large number of
cells in the body, in the early/mid-1980s, Eric Newsholme was able to advance evidence that glutamine
was an important modulator of leukocyte function, such as in lymphocytes [7] and macrophages [88].
One of the authors of this review, Newsholme P et al. (1986; 1987) [88–90], reported for the first
time that macrophages utilize glutamine actively. Pithon-Curi et al., in 1997 [91,92] described for the
first time the consumption of glutamine by neutrophils. The studies by Eric and Philip Newsholme
on glutamine metabolism in lymphocytes and macrophages, respectively, prompted many other
publications, which jumped from an average of two or three publications per year in the late 1960s and
early 1970s to about 50 publications per year in the last 20 years.
During infection and/or high catabolism, the rate of glutamine consumption by all immune
cells is similar or greater than glucose [89,90]. However, the increased demand for glutamine by
immune system cells, along with the increased use of this amino acid by other tissues, such as the
liver, may lead to a glutamine deficit in the human body. In addition, one of the most important
sites of glutamine synthesis, the skeletal muscles, reduce their contribution to maintaining plasma
glutamine concentration (Figure 2). This effect, depending on the situation may significantly contribute
to worsening diseases and infections, and/or increase the risk of subsequent infection, with possible
life-threatening implications [93].
In immune cells, glucose is mainly converted into lactate (glycolysis), whereas glutamine is
converted into glutamate, aspartate, and alanine by undergoing partial oxidation to CO2 , in a
process called glutaminolysis [3] (Figure 3). This unique conversion plays a key role in the effective
functioning of immune system cells. Furthermore, through the pentose phosphate pathway, a metabolic
pathway parallel to the glycolysis pathway, cells can produce ribose-5-phosphate (a five-carbon
sugar), which is a precursor for the pentose sugars seen in the RNA and DNA structure, as well as
glycerol-3-phosphate for phospholipid synthesis [94]. On the other hand, the degradation of glutamine,
and thus formation of NH3 , and aspartate leads to the synthesis of purines and pyrimidines of the DNA
and RNA. The expression of several genes in immune system cells is largely dependent on glutamine
availability [3]. For example, the role glutamine plays in the control of proliferation of immune system
cells occurs through activation of proteins, such as ERK and JNK kinases. Both proteins act on the
activation of transcription factors, such as JNK and AP-1, and it leads to the transcription of cell
proliferation-related genes. For instance, appropriate glutamine concentration leads to the expression
of key lymphocyte cells surface markers, such as CD25, CD45RO, and CD71, and the production of
cytokines, such as interferon-gamma (IFN-γ), TNF-α, and IL-6 [2,31,95,96]. Thus, glutamine acts as an
energy substrate for leukocytes and plays an essential role in cell proliferation, tissue repair process
activity, and intracellular pathways associated with pathogen recognition [97].
4.1. Neutrophils
The primary substrate for neutrophil survival endocytosis and ROS generation is glucose.
However, glucose is not the only energy metabolite source by these cells. Interestingly, when compared
to other leukocytes, such as macrophages and lymphocytes, neutrophils consume glutamine at the
highest rates [98,99]. Much of the glutamine is converted to glutamate, aspartate (via Krebs cycle
activity), and lactate in neutrophils. Under appropriate conditions, CO2 , glutamine, and glutamate
play an important role in the generation of essential compounds for leukocytes’ metabolism and
function, including GSH. Neutrophils use protein structures composed of uncondensed chromatin
and of antimicrobial factors also called neutrophilic extracellular traps (NETs). The action of NETs
requires ROS formation, synthesis of enzymes, such as myeloperoxidase (MPO) and elastase, as well
as components capable of overriding virulence factors and destroying extracellular bacteria [100].
The process involving ROS depends on the activation of the NADPH oxidase 2 (NOX2) complex.
161
Nutrients 2018, 10, 1564
Based on glutamine, the malate synthesis uses malic enzyme to produce substantial amounts of
NADPH, since it is necessary to form the superoxide anion (O2 − ), which presents antimicrobial activity.
Similarly, macrophages use glutamine for arginine and thus nitric oxide (NO) synthesis through the
action of the inducible NO synthase (iNOS) enzyme, by using NADPH as an energy source. Glutamine
increases superoxide generation through NADPH oxidase in neutrophils. 6-Diazo-5-oxo-L-norleucine
(DON), an inhibitor of phosphate-dependent glutaminase and thus of glutamine metabolism, causes a
significant decrease in superoxide production by neutrophils stimulated with phorbol myristate acetate
(PMA). PMA raises mRNA’s expression of gp91, p22, and p47, major components of the NADPH
oxidase complex. Glutamine increases expressions of these three proteins either in the absence
or in the presence of PMA. Glutamine enhances superoxide production in neutrophils, probably
via the generation of ATP and regulation of the expression of components of the NADPH oxidase
complex [101]. Glutamine plays a role to prevent the changes in NADPH oxidase activity and
superoxide production induced by adrenaline in neutrophils [102].
4.2. Macrophages
Metabolism of glucose and glutamine is profoundly affected during the macrophage
activation process [103,104]. The effects of thioglycollate (an inflammatory stimulus) and Bacillus
Calmette-Guérin—BCG (an activation stimulus) on macrophage glucose and glutamine metabolism
have been studied [105]. Either thioglycollate or BCG enhances activities of hexokinase and citrate
synthase, and also glucose oxidation whereas BCG markedly increases glutamine metabolism.
Lipopolysaccharide (LPS) administration also causes pronounced changes in macrophage metabolism
and function (for a review, see Nagy and Haschemi [106]. Glucose and glutamine metabolism is
also involved in polarizing signals that up-regulate the transcriptional programs required in the
macrophage capacity to perform specialized functions. Protein kinase B (PKB or Akt), mTOR complex
1 (mTORC1), mTORC2, and AMP-activated protein kinase (AMPK) play a critical role in metabolic
pathways and associated signalling activation [107,108]. For instance, extracellular glutamine may
function as the specific starvation-induced nutrient signal to regulate mTORC1. [17]. Synthesis and
secretion of pro-inflammatory cytokines, such as TNF-α, IL-1, and IL-6, by macrophages are also
regulated by glutamine availability.
Different populations of macrophages have now been identified, such as M1 and M2 [109–111].
The M1 and M2 are in fact two extremes of a still not completely known spectrum of macrophage
activation states [109,111,112]. Reprogramming signalling pathways are involved in the formation
of M1 or M2 phenotype macrophages. The metabolic reprogramming of macrophages include
key changes in glutamine and glucose metabolism [113]. No reports identified the requirement
of fatty acids for human macrophage IL-4 induced polarization [114]. This issue, however, remains
controversial. Interestingly, macrophages reprogram their metabolism and function to polarize for
pro- or anti-inflammatory cells, and this is a consequence of the environmental conditions and
stimuli [115]. Treatment of macrophages with LPS promotes a switch from glucose-dependent
oxidative phosphorylation to aerobic glycolysis—the Warburg effect [116]. Pyruvate kinase M2
regulates the hypoxia-inducible factor 1-alpha (Hif-1α) activity and IL-1β expression, being a key
molecule to induce the Warburg effect in LPS-activated macrophages [117]. Due to this mechanism,
M1 macrophages exhibit a quick increase in ATP formation that is required for the host defence
response [113,118,119]. The TCA cycle of M2 macrophages has no metabolic flux escape whereas
M1 macrophages (treated with LPS) have two points of substrate flux deviation, one occurring
at the isocitrate dehydrogenase step reaction and another one at post succinate formation. As a
result, there is an accumulation of TCA cycle intermediates (e.g., succinate, α-ketoglutarate, citrate,
and itaconate) that regulates LPS macrophage activation [119]. Itaconate has anti-inflammatory
properties through activation of nuclear factor erythroid 2-related factor 2 (Nrf2) via Kelch-like
ECH-associated protein 1 (KEAP1) alkylation [97]. The glutamine seems to be fully required for
IL-4 induction of macrophage alternative activation [120,121]. Liu, et al. [122] reported α-ketoglutarate,
162
Nutrients 2018, 10, 1564
4.3. Lymphocytes
Lymphocyte activation is associated with specific metabolic pathways to optimize its function.
The integration of multiple extracellular signals affects transcriptional programs and signalling
pathways that determine, in CD4+ T cells, for example, multiple events that include modulation
of energy metabolism, cell proliferation, and cytokine production. Associated bioenergetic processes
are dependent on the activation of AMPK, indicating cross-talk between metabolism and signalling
pathways in immune cell differentiation. Greiner, et al. [124] reported in rat thymocytes that
the use of the anaerobic glycolytic pathway is strongly increased after antigenic stimulation with
Concanavalin (ConA). Eric Newsholme’s group was the first to report the utilization of glutamine
by lymphocytes [125]. Glutamine plays an important role for the function of these cells in
different ways. Pyruvate is a common product of glucose and glutamine metabolism in the cells.
Curi, et al. [126] described that mesenteric lymphocytes have increased pyruvate oxidation through
pyruvate carboxylase when stimulated with ConA, indicating that both glucose and glutamine are
involved in the control of lymphocyte proliferation and function. Mitochondria have been reported to
be able to regulate leukocyte activation. Succinate, fumarate, and citrate, metabolites of the Krebs cycle
and produced through glucose and glutamine metabolism, participate in the control of immunity and
inflammation either in innate and adaptive immune cells [97].
Most glucose molecules are transported via glucose transporter 1 (GLUT1), which is not observed
in non-activated lymphocytes [127]. GLUT1 is an important metabolic marker of lymphocyte activation
as it migrates rapidly to the cell surface after stimulation. Glucose deprivation causes a lower rate
of basal proliferation, as well as increased production of IL-2, TNF-α, INF-γ, and IL-4 by CD4+
T cells [128]. Activation of intracellular signalling by Akt beyond GLUT1 protein levels further
increases glucose uptake and T cell activation. mTOR and AMPK play important and distinct roles
in metabolism and immunity. The stimulation of lymphoid cells leads to increased GLUT1 uptake
of glucose by acting on mTOR protein [129]. This pathway is also involved in the differentiation
of CD4+ T-cell subsets since mTOR-deficient mice have a decrease in differentiation for effector T
lymphocytes [130,131]. In contrast, the AMPK pathway inhibits mTOR by suppressing the signalling of
this protein and promotes activation of mitochondrial oxidative metabolism rather than the glycolytic
pathway [132,133]. Glutamine is required for T and B-lymphocytes’ proliferation process, as well as for
protein synthesis, IL-2 production, and antibody synthesis rates presented by these cells. The evidence
has now accumulated that glutamine metabolism plays a key role in the activation of lymphocytes.
Glutamine is required for human B lymphocyte differentiation to plasma cell and to lymphoblastic
transformation [134].
The cell proliferation process requires both ATP for high-energy expenditure and precursors for
the biosynthesis of complex molecules, such as lipids (cholesterol and triglycerides) and nucleotides
for RNA and DNA synthesis. To perform rapid proliferation activity under the certain stimulus,
lymphocytes switch from oxidative phosphorylation to aerobic glycolysis plus glutaminolysis, and so
markedly increase glucose and glutamine utilization. The metabolic transition in a Th0 lymphocyte is
crucial for the activation of T cells, since glucose metabolism provides intermediates for the biosynthetic
pathways, being a prerequisite for the growth and differentiation of T cells [135]. Glycolysis plays an
important role in effector T cell functions associated with the production of inflammatory cytokines,
163
Nutrients 2018, 10, 1564
mainly INF-γ and IL-2 [136]. The blockade of glyceraldehyde 3-phosphate dehydrogenase (GAPDH)
mRNA by the use of siRNA promotes a reduction of INF-γ in lymphocytes [136]. Therefore, the high
glycolytic activity is closely associated with the differentiation of Th0 to Th1 cells [132]. Inhibition of
the glycolytic pathway blocks this process whereas it promotes differentiation into Treg cells. Increased
glycolysis by proliferating cells is linked to increased uptake of glucose and increased expression and
activity of glycolytic enzymes, whereas glucose utilization in the oxidative phosphorylation pathway
(OXPHOS) is decreased. Therefore, the “metabolic switch” meets the higher energy requirements,
generates metabolic intermediates required for the biosynthesis of macromolecules, and suppresses the
metabolic features of rest lymphocytes. Inadequate nutrient delivery or specific metabolic inhibition
prevents the activation and proliferation of T cells since the inability to use glucose inhibits T cell
differentiation in vitro and in vivo [135,137]. Mitochondrial dynamics are closely associated with
T lymphocyte metabolism and function. Activated effector T cells have punctate mitochondria
and augmented the activity of anabolic pathways whereas memory T lymphocytes exhibit fused
mitochondria and enhanced oxidative phosphorylation activity [138]. HIF1-α plays a central role in
the maturation of dendritic cells and the activation of T cells. This factor controls leukocyte metabolism
reprogramming, through changes in gene expression, and thus immune cell functions [139].
Glycolysis and glutaminolysis are strongly associated to ensure appropriateness for lymphocyte
function. Hexosamine biosynthesis requires glucose and glutamine for the de novo synthesis of
uridine diphosphate N-acetylglucosamine (UDP-GlcNAc). This sugar-nucleotide inhibits receptor
endocytosis and signalling through promoting N-acetylglucosamine branching of Asn (N)-linked
glycans. Araujo, et al. [140] reported that high aerobic glycolysis and glutaminolysis activities in a
co-operative way decrease the UDP-GlcNAc synthesis and N-glycan branching in mouse T cell blasts
due to the low availability of these metabolites for hexosamine synthesis. As a consequence, growth
and pro-inflammatory TH 17 features prevail over anti-inflammatory-induced T regulatory (iTreg)
differentiation. The latter process is promoted by IL-2 receptor-α (CD25) loss through endocytosis.
The authors then postulated that a primary function of concomitant high aerobic glycolysis and
glutaminolysis activities is to limit precursors to N-glycan biosynthesis. This metabolic feature of
T lymphocytes has marked implications in autoimmunity and cancer. Glutamine also serves as a
precursor for the synthesis of putrescine and the polyamines, spermidine and spermine. High levels of
polyamines are reported in tumour cells and in autoreactive B- and T-cells in autoimmune diseases.
Polyamines have been described to play a role in the control of normal immune cell function and have
been associated with autoimmunity and anti-tumour immune cell properties [141].
164
Nutrients 2018, 10, 1564
intracellular signalling pathways [2]. Glutamine action also involves signalling pathways’ activation
by phosphorylation, such as NF-κB and MAPKs [144]. Thus, the function of glutamine goes beyond
that of a metabolic fuel or protein synthesis precursor. This amino acid is also an important regulator
of leukocyte function, acting on either gene expression or signalling pathways’ activation.
165
Nutrients 2018, 10, 1564
proteins clustered according to their molecular weight, which have many intracellular functions.
Possibly, the most important function displayed by HSP’s are the action of a molecular chaperone.
This function assists protein transport, prevents protein aggregation during folding, and protects newly
synthesized polypeptide chains against misfolding and protein denaturation [151]. Although several
HSP families have been studied in the last couple of years (e.g., HSP10, HSP25, HSP27; HSP90), the most
famous and well described in the literature is the HSP70 (i.e., HSP72 + HSP73) family [29,151,152].
HSP70 acts as anti-inflammatory protein by virtue of turning NF-κB off and attenuating the production
of inflammatory mediators [153]. Moreover, HSP70 modulate autophagy by regulating the mTOR/Akt
pathway and block signalling pathways associated with protein-degradation [152].
Glutamine found at concentrations similar to those recorded for human plasma leads to a
significant HSP72 gene expression increase in peripheral blood mononuclear cells subjected to LPS
treatment. On the other hand, reduced glutamine concentration results in reduced HSP72 expression
in monocytes; this effect depends on mRNA stability. The preoperative administration of glutamine
can modulate HSP70 expression by reducing the activation levels of the cyclic AMP response element
binding protein (CREB), which is often associated with exacerbated inflammatory responses. This effect
depends on the iNOS activity and leads to an NO production increase. Other studies corroborated the
present results and presented similar mechanisms, as well as effects on the expression of other HSP’s,
such as HSP25, HSP27, and HSP90.
Glutamine plays a crucial role in the modulation of HSP’s expression through the
hexosamine biosynthetic pathway (HBP, Figure 3) [21,94]. In the HBP, glutamine leads to the
production of UDP-GlcNAc and (UDP)-N-acetylgalactosamine (UDP-GalNAc) through the enzyme,
fructose-6-phosphate amidotransferase (GFAT, the first and rate-limiting step of HBP). UDP-GlcNAc
and UDP-GalNAc, in turn, may be attached to serine or threonine hydroxyl moieties in nuclear
and cytoplasmic proteins by the enzymic action of O-linked-N-acetylglucosaminyl (O-GlcNAc)
transferase (a.k.a. OGT) [94]. The main donors for UDP-GlcNAc are glucose, glutamine, and uridine
triphosphate (UTP) from the HBP. Interestingly, both nutrients’ availability and cell stress affect
O-GlcNAc downstream signalling, and, not surprisingly, this mechanism is also altered in several
metabolic diseases, infection, and inflammatory processes [154,155]. O-GlcNAc synthesis leads to the
activation of many transcriptional factors, for instance, Sp1, phosphorylation of Eukaryotic Initiation
Factor 2 (eIF2), and sirtuin-1 (SIRT1) [156]. Both Sp1 and eIF2 are key transcription factors for the
induction of the main thermal shock eukaryotic factor, HSF-1 [157]. Alternatively, SIRT1 enhances
HSF-1 expression and prolongs its activation by binding to the promoters of HS genes, leading to the
HSP’s expression [94,158]. Although the O-GlcNAc/Sp1 pathway is considered the main mechanism
of the HSP’s gene expression and production, glutamine may also act on HBP via p38/MAPK,
leading to the HSP’s expression in cells, such as neutrophils. This response may explain the reduction
in neutrophils’ apoptosis after high-intensity physical exercise [144]. Furthermore, by increasing the
HBP flux, glutamine stimulates the HSP’s expression by blocking the glycogen synthase kinase 3 beta
(GSK-3β), an enzyme that constitutively inhibits HSF-1 activation by phosphorylating the transcription
factor at Ser303 [94,159].
In vitro [21,160] and in vivo [4,39,161–163] studies demonstrate that glutamine availability
maintains cell homeostasis and promotes cell survival against environmental and physiological
stress challenges through an enhanced protection mediated by intracellular HSP (iHSP) levels [94].
Interestingly, under severe infection and/or catabolism, low glutamine availability in the body can
eventually be accompanied by an aberrant iHSP and lead to the HSP’s release to the extracellular
space (eHSP) [4]. eHSP have a wide variety of effects on other cells, including impacting on a cell to
cell interaction and chemotaxis, and in some cases, act as a signal to the immune and inflammatory
responses. On the other hand, eHSP can also function as a stress signalling and pro-inflammatory
molecule by interacting with Toll-like receptors 2 (TLR2) and 4 (TLR4) [164]. This effect can
down-regulate iHSP in many cells, leading to apoptosis [4], and has also been associated with increased
insulin resistance in skeletal muscle cells [165], and β-cell failure in diabetic individuals. Currently, a
166
Nutrients 2018, 10, 1564
novel and overall index of immunoinflammatory status, the extracellular to intracellular HSP70 ratio
index (H-index), measured in peripheral blood mononuclear cells (PBMCs) [94], has been established.
Table 1. Total protein, glutamine, glutamate, and leucine (g/100g food) content in some animal and
vegetable foods using the gene sequencing method (adapted from [166]).
On the contrary, during major and/or critical illness, sepsis, trauma, and post-surgery
circumstances, patients suffer from chronic weakness and several nutritional limitations (e.g., state of
unconsciousness, gastrointestinal disturbances, and/or chew related problems), which impair
homeostasis, and are associated with poor clinical outcomes. Severe disturbances in amino
acid metabolism and/or intermediary metabolism followed by skeletal muscle proteolysis are
key characteristics of hypermetabolic/hypercatabolic states [167]. During hypercatabolism, some
non-essential amino acids, including glutamine, become conditionally essential. As previously
mentioned, glutamine is critical for cell homeostasis, and cells cannot survive and/or proliferate in an
environment where glutamine is lacking. Therefore, the administration of non-synthetic amino acid
supplements, such as glutamine, has been a research target in in the last many years and is currently
indicated for hypercatabolic and/or ill patients. However, the efficacy of glutamine supplementation
is frequently questioned due to confusing and controversial results [150,168,169].
Glutamine is usually administrated by utilizing its free form (also known as an isolated
amino acid), or bond with another amino acid, also known as the dipeptide form (Figure 4).
Several glutamine dipeptides with potential recovery health benefits have been described, such as
L -glycyl- L-glutamine (Gly-Gln) and L -arginyl- L -glutamine (Arg-Gln); however, the most well-known
is possibly L-alanyl-L-glutamine (Ala-Gln) [170]. Given parenterally, many clinical and experimental
studies and appropriate systematic reviews [168] have concluded that glutamine dipeptides can reduce
the rate of infectious complications [171–174], length of hospital stay [9,175], and mortality of critically
ill patients [10,176,177]. The choice for free glutamine or glutamine dipeptides largely depends on the
patient’s catabolic circumstance and/or the most suitable route of administration (e.g., enteral and
parenteral nutrition). For instance, in patients receiving total parenteral nutrition (TPN), glutamine
dipeptides offer several advantages, such as stability during sterilization, prolonged storage, and high
range of solubility (154 g/L H2 O at 20 ◦ C, 568 g/L H2 O at 20 ◦ C, respectively) when compared to free
glutamine (36 g/L H2 O at 20 ◦ C) [170]. Moreover, free glutamine is usually commercially available as a
crystalline amino acid powder and can be diluted into commercially available TPN solutions, however,
167
Nutrients 2018, 10, 1564
this procedure requires daily preparations at a controlled temperature (i.e., 4 ◦ C), aseptic conditions
followed by sterilization through specific membrane filtration, and the concentration should not exceed
1–2%. This is particularly important because, at a low concentration, TPN solutions will increase the
patient’s fluid intake to meet the daily glutamine recommendation, however, this cannot be feasible
for fluid restricted patients.
Figure 4. Mechanisms of enteral and parenteral glutamine (GLN) supply. Glutamine is an important
substrate for rapidly dividing cells, such as enterocytes. This is a major site of glutamine consumption
obtained from both exogenous/diet (luminal membrane) and/or endogenous glutamine synthesis
(basolateral membrane). Free glutamine supplementation is mainly metabolized in the gut and poorly
contribute to glutaminemia and tissue stores. On the other hand, glutamine dipeptides (e.g., Ala-Gln,
Gly-Gln, Arg-Gln) escape from the gut metabolization and quickly supply glutamine to the plasma
and target tissues. This effect is mainly attributed to the oligopeptide transporter 1 (Pept-1) located in
the luminal membrane of the enterocytes.
In both oral/enteral or parenteral nutrition, the typical glutamine daily administration (free
and dipeptide forms) may vary from a fixed dose of 20–35 g/24 h to an adjusted dose of <1.0 g
(usually 0.3 g–0.5 g) per kg of body weight [168]. As any other nutrient or medication administrated
directly into the bloodstream when free or dipeptide forms of glutamine are given parenterally,
the increase in plasma glutamine is superior when compared to oral/enteral feeds [178]. However,
it should be noted that although parenteral routes can secure nutrient delivery to target tissues, it is
always an invasive route and may increase the risk of infections per se. It is strongly recommended
that the decision to use parenteral solutions must be based on several nutritional parameters, such as
poor nutritional status, dramatic reduction of body weight and body mass index, low plasma albumin,
and/or severe loss of nitrogen and tissue function.
For individuals with regular enteral feeds at home or hospitals, and also elite athletes
where glutamine supplementation is eventually recommended, oral or enteral routes are always
more physiological. Furthermore, enteral solutions stimulate the intestinal cells to produce other
intermediary amino acid derivatives important for immunological functions, and also compromised
168
Nutrients 2018, 10, 1564
in hypercatabolic patients, such as arginine and its downstream metabolites (e.g., ornithine and
citrulline) [179]. Experimental studies in animal models and humans have shown an increase
in glutaminemia between ±30 to ±120 min after oral/enteral free glutamine [50] or Ala-Gln
supplementation [50,177]. However, the peak concentration and the area under the curve promoted
by Ala-Gln tend to be superior when compared to free glutamine supply. This effect is largely due
to the expression of the human oligopeptide transporter 1 (Pept-1) located in the luminal microvilli
membrane of the enterocytes [180], and in a lesser extent, through paracellular mechanisms and
cell-penetrating peptides’ translocation [181,182] (Figure 4). Pept-1 is a high capacity, low-affinity
proton-coupled cotransporter of diverse di/tripeptides, which include the glutamine dipeptides.
Pept-1 is considered the main route of protein absorption in mammals’ intestines since the protein
is able to transport about 400 dipeptides and 8000 tripeptides derived from the 20 L-α amino
acids [180]. As a result, free glutamine and/or alanine deriving from enteral Ala-Gln administration
can be released into the bloodstream, thus making the amino acids available to target tissues,
including the liver [183], immune system [39], kidneys [184] and skeletal muscles [53] (Figure 4).
Interestingly, the effects promoted by Ala-Gln are also mediated by the presence of the amino acid,
alanine, in the peptide formulation. Oral free glutamine along with free alanine promoted similar
metabolic, antioxidant, and immunological effects when compared to Ala-Gln supplementation in
in vivo animal models submitted to infection [4,39], and exhaustive aerobic [30,45,53] and resistance
physical exercise [29,162]. Importantly, in all of these experiments, the supplemented groups received
isonitrogenous and isocaloric solutions, i.e., both containing 13.46 g of glutamine/100 mL and 8.20 g
of alanine/100 mL. Although the precise mechanisms are still unknown, it is clear that both amino
acids work in parallel, especially in absorptive cells (Figure 4). For instance, alanine is rapidly
metabolized via alanine aminotransferase to pyruvate, with concomitant production of glutamate
from 2-oxoglutarate, which contribute to antioxidant defence mediated by GSH. Although other free
amino acid combinations need to be tested, these important discoveries may lead to the design of new
formulations for specific hypercatabolic patients.
Oral/enteral or parenteral doses of glutamine supplementation have been tested in hundreds of
studies in both animal models and humans, and if offered as a single nutrient supplementation
(not combined with other additives), can be considered safe. In addition, there is no scientific
evidence demonstrating that glutamine supplementation can suppress and/or inhibit permanently its
endogenous production or de novo synthesis. However, as any other amino acid offered in excessive
doses, it can promote hyperaminoacidemia and result in poor clinical outcomes. It is considered not to
be the best practice to provide glutamine supplementation to patients without a proper evaluation
that is supported by a nutritional assessment and biochemical laboratory tests.
Glutamine metabolism and supplementation in cancer has also raised many concerns among
the scientific community, and deserve some comments. It is well established that cancer cells are
extremely dependent of glutamine metabolism and availability, however, the role played by glutamine
in cancer/tumours cells in vivo is still controversial, and thus the effects of the supplementation.
Cancer cells take advantage of aerobic glycolysis (also known as the Warburg effect), and therefore
glucose to maintain the supraphysiological survival and growth [185,186]. On the other hand, there is
an increasing evidence of the role of oncogenes and tumour suppressors in the regulation of nutrient
metabolism [187]. Aberrant mutations in these genes lead to altered nutrient metabolism, and can
significantly contribute to the development and/or progression of cancer cells [186]. For instance,
glucose, glutamine, lipids, and acetate can be utilized as carbon and energy sources [25]. To increase
the level of complexity, the nutrient sources may vary among different types of cancer and/or tumour
cells and are highly heterogeneous [25,187]. For example, lung cancer cell lines are highly dependent
of glutamine supply in vitro, however, in vivo experiments demonstrate that glucose is the preferred
source of carbons supplied to the Krebs cycle, through the action of pyruvate carboxylase [188],
with little changes in glutamine consumption [187]. Human and mouse gliomas exhibit high
rates of glucose catabolism, and use glucose to synthesize glutamine through glutamate-ammonia
169
Nutrients 2018, 10, 1564
ligase (GLUL), which in turn promotes nucleotide biosynthesis via the pentose phosphate pathway
independently of circulating glutamine [189,190]. Conversely, prostate cancer cell lines exhibit aberrant
intracellular lipid metabolism [191], and an increased gene expression of glutaminolitic enzymes and
glutamine transporters, thereby stimulating cell growth via glutamine uptake [192].
The variances in cancer cells’ nutrient metabolism suggest that different nutritional approaches
should be taken into consideration. However, studies have also targeted whether a glutamine
exogenous supply may attenuate the side effects promoted by chemotherapy and radiation in cancer
patients [193]. In a systematic review, Sayles, et al. [194] reported that in 11 of 15 studies, oral glutamine
supplementation (dose range: 30 g/day in 3 divided doses, or 7.5–24 g/day) significantly reduced the
grade of mucositis (common in 90% of head/neck cancer patients) [193] and/or attenuated weight loss
in cancer patients. In a double-blind, placebo-controlled, randomized trial, colorectal cancer patients
undergoing chemotherapy were supplemented with 18 g/day of glutamine (five days before and
during the treatment). Glutamine treatment reduced the side-effects induced by chemotherapy, such
as intestinal absorption and permeability, diarrhea, and gut mucositis [195]. As a whole, it is important
to highlight that glutamine supplementation for cancer patients may also fuel certain types of cancer
cells, and have a negative impact on health. However, considering the strong ability of metabolic
switching of cancer cells, glucose and/or lipids can also induce similar effects, and therefore it would
be difficult for a human to survive and maintain immunity and/or immune surveillance without
these nutrients. As mentioned previously, a proper and possibly individualized patient evaluation is
required to determine the suitability of glutamine supplementation.
Low plasma glutamine level (hypoglutaminemia) is usually used as a parameter to indicate
the need for a glutamine exogenous supply. However, the correlation between the concentration of
glutamine in plasma and tissue vary significantly between hypercatabolic patients, and therefore
among studies [150]. For instance, muscle glutamine was dramatically reduced in abdominal surgery
patients, but no changes were detected in plasma [196]. In critically ill patients, however, there is
a profound drop in muscle glutamine, but a variable reduction in plasma [150]. Other important
glutamine sites, such as the gut and the liver, may show a concomitant plasma and tissue glutamine
reduction, or even an inverse relationship during major illness [86,197,198]. These findings are also
in agreement with data obtained in rats [29,45,183] and mice [4,199] submitted to infection and
exhaustive exercise.
The variations between plasma and tissue glutamine concentration are due to the fact that only
a small fraction of the total body free glutamine is in plasma. To add to the confusion, it is known
that cells of the immune system, such as lymphocytes maintained in a low glutamine availability
(similar to low plasma glutamine concentration, e.g., ±400 μM), still proliferate, when compared to
resting values (e.g., 600 μM) [4,7,90]. The rate of macrophages’ phagocytosis and cytokine production
by diverse peripheral blood mononuclear cells (PBMCs) is also dependent on glutamine availability,
with decreasing rates at ±<600 μM of glutamine [21,200]. Thus, for some catabolic/hypercatabolic
patients, the changes in glutamine concentration, especially in plasma, will not necessarily affect
and suppress immune functions, and possibly no significant changes might be observed in data
obtained from immune parameters and function, and mortality risk predictors, such as APACHE
II or SAPS III. Taken together, hypoglutaminemia can only be interpreted as an independent
variable of mortality and/or poor clinical outcome [10,93,168,169,174,201]. More in-depth studies
are required to explore the specific relationship between dramatic changes in plasma glutamine and
outcomes in critically ill patients. Currently, the decision of glutamine supplementation should be
based in a set of immune-inflammatory parameters allied with appropriate nutritional assessment,
and eventually, risk predictors. In addition, glutamine supplementation studies cannot be judged
on trials where only the very sickest patients (i.e., with two or more organ failures) were eligible for
this nutritional intervention, and in these situations, supraphysiologic doses are not an appropriate
nutritional solution.
170
Nutrients 2018, 10, 1564
Author Contributions: V.C. conceived the idea, and with the help of M.M.R. organised and designed the
manuscript. V.C., M.M.R., K.N.K. and R.C. wrote the first draft, which was reviewed and revised by P.N. Figures
and Table 1 were designed by V.C. All authors contributed significantly in the manuscript revision and agreed
with the final submitted version.
Funding: We thank the Department of Health Science, Torrens University, and the Curtin School of Pharmacy
and Biomedical Sciences for financial research support and excellent research facilities, respectively.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Grohmann, U.; Mondanelli, G.; Belladonna, M.L.; Orabona, C.; Pallotta, M.T.; Iacono, A.; Puccetti, P.; Volpi, C.
Amino-acid sensing and degrading pathways in immune regulation. Cytokine Growth Factor Rev. 2017, 35,
37–45. [CrossRef] [PubMed]
2. Curi, R.; Lagranha, C.J.; Doi, S.Q.; Sellitti, D.F.; Procopio, J.; Pithon-Curi, T.C.; Corless, M.; Newsholme, P.
Molecular mechanisms of glutamine action. J. Cell. Physiol. 2005, 204, 392–401. [CrossRef] [PubMed]
3. Curi, R.; Newsholme, P.; Marzuca-Nassr, G.N.; Takahashi, H.K.; Hirabara, S.M.; Cruzat, V.; Krause, M.;
de Bittencourt, P.I.H., Jr. Regulatory principles in metabolism-then and now. Biochem. J. 2016, 473, 1845–1857.
[CrossRef] [PubMed]
4. Cruzat, V.F.; Pantaleao, L.C.; Donato, J., Jr.; de Bittencourt, P.I.H., Jr.; Tirapegui, J. Oral supplementations
with free and dipeptide forms of l-glutamine in endotoxemic mice: Effects on muscle glutamine-glutathione
axis and heat shock proteins. J. Nutr. Biochem. 2014, 25, 345–352. [CrossRef] [PubMed]
5. Newsholme, P. Why is l-glutamine metabolism important to cells of the immune system in health, postinjury,
surgery or infection? J. Nutr. 2001, 131, 2514S–2523S. [CrossRef] [PubMed]
6. Cruzat, V.F.; Krause, M.; Newsholme, P. Amino acid supplementation and impact on immune function in
the context of exercise. J. Int. Soc. Sports Nutr. 2014, 11, 61. [CrossRef] [PubMed]
7. Ardawi, M.S.M.; Newsholme, E.A. Maximum activities of some enzymes of glycolysis, the tricarboxylic acid
cycle and ketone-body and glutamine utilization pathways in lymphocytes of the rat. Biochem. J. 1982, 208,
743–748. [CrossRef] [PubMed]
8. Flaring, U.B.; Rooyackers, O.E.; Wernerman, J.; Hammarqvist, F. Glutamine attenuates post-traumatic
glutathione depletion in human muscle. Clin. Sci. 2003, 104, 275–282. [CrossRef] [PubMed]
9. Roth, E. Nonnutritive effects of glutamine. J. Nutr. 2008, 138, 2025S–2031S. [CrossRef] [PubMed]
171
Nutrients 2018, 10, 1564
10. Rodas, P.C.; Rooyackers, O.; Hebert, C.; Norberg, A.; Wernerman, J. Glutamine and glutathione at icu
admission in relation to outcome. Clin. Sci. 2012, 122, 591–597. [CrossRef] [PubMed]
11. Newsholme, E.A.; Parry-Billings, M. Properties of glutamine release from muscle and its importance for the
immune system. J. Parenter. Enter. Nutr. 1990, 14, 63S–67S. [CrossRef] [PubMed]
12. Wernerman, J. Clinical use of glutamine supplementation. J. Nutr. 2008, 138, 2040S–2044S. [CrossRef]
[PubMed]
13. Berg, A.; Norberg, A.; Martling, C.R.; Gamrin, L.; Rooyackers, O.; Wernerman, J. Glutamine kinetics
during intravenous glutamine supplementation in icu patients on continuous renal replacement therapy.
Intensive Care Med. 2007, 33, 660–666. [CrossRef] [PubMed]
14. Labow, B.I.; Souba, W.W.; Abcouwer, S.F. Mechanisms governing the expression of the enzymes of glutamine
metabolism—Glutaminase and glutamine synthetase. J. Nutr. 2001, 131, 2467S–2486S. [CrossRef] [PubMed]
15. Cruzat, V.F.; Newsholme, P. An introduction to glutamine metabolism. In Glutamine; CRC Press: Boca Raton,
FL, USA, 2017; pp. 1–18.
16. Cooney, G.; Curi, R.; Mitchelson, A.; Newsholme, P.; Simpson, M.; Newsholme, E.A. Activities of some key
enzymes of carbohydrate, ketone-body, adenosine and glutamine-metabolism in liver, and brown and white
adipose tissues of the rat. Biochem. Biophys. Res. Commun. 1986, 138, 687–692. [CrossRef]
17. Tan, H.W.S.; Sim, A.Y.L.; Long, Y.C. Glutamine metabolism regulates autophagy-dependent mtorc1
reactivation during amino acid starvation. Nat. Commun. 2017, 8, 338. [CrossRef] [PubMed]
18. Ardawi, M.S. Glutamine metabolism in the lungs of glucocorticoid-treated rats. Clin. Sci. 1991, 81, 37–42.
[CrossRef] [PubMed]
19. Parry-Billings, M.; Dimitriadis, G.D.; Leighton, B.; Bond, J.; Bevan, S.J.; Opara, E.; Newsholme, E.A. Effects
of hyperthyroidism and hypothyroidism on glutamine metabolism by skeletal muscle of the rat. Biochem. J.
1990, 272, 319–322. [CrossRef] [PubMed]
20. Parry-Billings, M.; Dimitriadis, G.; Leighton, B.; Dunger, D.; Newsholme, E. The effects of growth hormone
administration in vivo on skeletal muscle glutamine metabolism of the rat. Horm. Metab. Res. 1993, 25,
292–293. [CrossRef] [PubMed]
21. Cruzat, V.F.; Keane, K.N.; Scheinpflug, A.L.; Cordeiro, R.; Soares, M.J.; Newsholme, P. Alanyl-glutamine
improves pancreatic beta-cell function following ex vivo inflammatory challenge. J. Endocrinol. 2015, 224,
261–271. [CrossRef] [PubMed]
22. Krebs, H.A. Metabolism of amino-acids: The synthesis of glutamine from glutamic acid and ammonia, and
the enzymic hydrolysis of glutamine in animal tissues. Biochem. J. 1935, 29, 1951–1969. [CrossRef] [PubMed]
23. Neu, J.; Shenoy, V.; Chakrabarti, R. Glutamine nutrition and metabolism: Where do we go from here?
FASEB J. 1996, 10, 829–837. [CrossRef] [PubMed]
24. Holecek, M. Branched-chain amino acids in health and disease: Metabolism, alterations in blood plasma,
and as supplements. Nutr. Metab. 2018, 15, 33. [CrossRef] [PubMed]
25. Altman, B.J.; Stine, Z.E.; Dang, C.V. From krebs to clinic: Glutamine metabolism to cancer therapy. Nat. Rev.
Cancer 2016, 16, 619–634. [CrossRef] [PubMed]
26. Kao, C.; Hsu, J.; Bandi, V.; Jahoor, F. Alterations in glutamine metabolism and its conversion to citrulline in
sepsis. Am. J. Physiol. Endocrinol. Metab. 2013, 304, E1359–E1364. [CrossRef] [PubMed]
27. Rogero, M.M.; Borges, M.C.; Pires, I.S.D.; Borelli, P.; Tirapegui, J. Ffect of glutamine supplementation
and in vivo infection with mycobacterium bovis (bacillus calmette-guerin) in the function of peritoneal
macrophages in early weaned mice. Ann. Nutr. Metab. 2007, 51, 173–174.
28. Karinch, A.M.; Pan, M.; Lin, C.M.; Strange, R.; Souba, W.W. Glutamine metabolism in sepsis and infection.
J. Nutr. 2001, 131, 2531S–2550S. [CrossRef] [PubMed]
29. Leite, J.S.; Raizel, R.; Hypolito, T.M.; Rosa, T.D.; Cruzat, V.F.; Tirapegui, J. L-glutamine and l-alanine
supplementation increase glutamine-glutathione axis and muscle hsp-27 in rats trained using a progressive
high-intensity resistance exercise. Appl. Physiol. Nutr. Metab. 2016, 41, 842–849. [CrossRef] [PubMed]
30. Cruzat, V.F.; Rogero, M.M.; Tirapegui, J. Effects of supplementation with free glutamine and the dipeptide
alanyl-glutamine on parameters of muscle damage and inflammation in rats submitted to prolonged exercise.
Cell Biochem. Funct. 2010, 28, 24–30. [CrossRef] [PubMed]
31. Curi, R.; Lagranha, C.J.; Doi, S.Q.; Sellitti, D.F.; Procopio, J.; Pithon-Curi, T.C. Glutamine-dependent changes
in gene expression and protein activity. Cell Biochem. Funct. 2005, 23, 77–84. [CrossRef] [PubMed]
172
Nutrients 2018, 10, 1564
32. Djoko, K.Y.; Phan, M.D.; Peters, K.M.; Walker, M.J.; Schembri, M.A.; McEwan, A.G. Interplay between
tolerance mechanisms to copper and acid stress in Escherichia coli. Proc. Nat. Acad. Sci. USA 2017, 114,
6818–6823. [CrossRef] [PubMed]
33. Wernerman, J. Feeding the gut: How, when and with what—The metabolic issue. Curr. Opin. Crit. Care 2014,
20, 196–201. [CrossRef] [PubMed]
34. Beutheu, S.; Ouelaa, W.; Guerin, C.; Belmonte, L.; Aziz, M.; Tennoune, N.; Bole-Feysot, C.; Galas, L.;
Dechelotte, P.; Coeffier, M. Glutamine supplementation, but not combined glutamine and arginine
supplementation, improves gut barrier function during chemotherapy-induced intestinal mucositis in
rats. Clin. Nutr. 2014, 33, 694–701. [CrossRef] [PubMed]
35. Souba, W.W.; Smith, R.J.; Wilmore, D.W. Glutamine metabolism by the intestinal tract. J. Parenter. Enter. Nutr.
1985, 9, 608–617. [CrossRef] [PubMed]
36. Holecek, M. Side effects of long-term glutamine supplementation. J. Parenter. Enter. Nutr. 2013, 37, 607–616.
[CrossRef] [PubMed]
37. Kim, M.H.; Kim, H. The roles of glutamine in the intestine and its implication in intestinal diseases. Int. J.
Mol. Sci. 2017, 18, 1051. [CrossRef] [PubMed]
38. Souba, W.W.; Herskowitz, K.; Salloum, R.M.; Chen, M.K.; Austgen, T.R. Gut glutamine metabolism. J. Parenter.
Enter. Nutr. 1990, 14, 45S–50S. [CrossRef] [PubMed]
39. Cruzat, V.F.; Bittencourt, A.; Scomazzon, S.P.; Leite, J.S.; de Bittencourt, P.I.H.; Tirapegui, J. Oral free and
dipeptide forms of glutamine supplementation attenuate oxidative stress and inflammation induced by
endotoxemia. Nutrition 2014, 30, 602–611. [CrossRef] [PubMed]
40. Aosasa, S.; Wells-Byrum, D.; Alexander, J.W.; Ogle, C.K. Influence of glutamine-supplemented caco-2 cells
on cytokine production of mononuclear cells. J. Parenter. Enter. Nutr. 2003, 27, 333–339. [CrossRef] [PubMed]
41. Coeffier, M.; Claeyssens, S.; Hecketsweiler, B.; Lavoinne, A.; Ducrotte, P.; Dechelotte, P. Enteral glutamine
stimulates protein synthesis and decreases ubiquitin mRNA level in human gut mucosa. Am. J. Physiol.
Gastrointest. Liver Physiol. 2003, 285, G266–G273. [CrossRef] [PubMed]
42. Jobin, C.; Hellerbrand, C.; Licato, L.L.; Brenner, D.A.; Sartor, R.B. Mediation by nf-kappa b of cytokine
induced expression of intercellular adhesion molecule 1 (icam-1) in an intestinal epithelial cell line, a process
blocked by proteasome inhibitors. Gut 1998, 42, 779–787. [CrossRef] [PubMed]
43. Coeffier, M.; Miralles-Barrachina, O.; Le Pessot, F.; Lalaude, O.; Daveau, M.; Lavoinne, A.; Lerebours, E.;
Dechelotte, P. Influence of glutamine on cytokine production by human gut in vitro. Cytokine 2001, 13,
148–154. [CrossRef] [PubMed]
44. Tirapegui, J.; Cruzat, V. Glutamine and skeletal muscle. In Glutamine in Clinical Nutrition; Rajendram, R.,
Preedy, V.R., Patel, V.B., Eds.; Springer: New York, NY, USA, 2015; pp. 499–511.
45. Cruzat, V.F.; Tirapegui, J. Effects of oral supplementation with glutamine and alanyl-glutamine on glutamine,
glutamate, and glutathione status in trained rats and subjected to long-duration exercise. Nutrition 2009, 25,
428–435. [CrossRef] [PubMed]
46. Walsh, N.P.; Blannin, A.K.; Robson, P.J.; Gleeson, M. Glutamine, exercise and immune function. Links and
possible mechanisms. Sports Med. 1998, 26, 177–191. [CrossRef] [PubMed]
47. Rowbottom, D.G.; Keast, D.; Morton, A.R. The emerging role of glutamine as an indicator of exercise stress
and overtraining. Sports Med. 1996, 21, 80–97. [CrossRef] [PubMed]
48. Curi, R.; Newsholme, P.; Procopio, J.; Lagranha, C.; Gorjao, R.; Pithon-Curi, T.C. Glutamine, gene expression,
and cell function. Front. Biosci. 2007, 12, 344–357. [CrossRef] [PubMed]
49. Rogero, M.M.; Tirapegui, J.; Pedrosa, R.G.; de Castro, I.A.; Pires, I.S.D. Effect of alanyl-glutamine
supplementation on plasma and tissue glutamine concentrations in rats submitted to exhaustive exercise.
Nutrition 2006, 22, 564–571. [CrossRef] [PubMed]
50. Rogero, M.M.; Tirapegui, J.; Pedrosa, R.G.; Pires, I.S.D.; de Castro, I.A. Plasma and tissue glutamine response
to acute and chronic supplementation with l-glutamine and l-alanyl-l-glutamine in rats. Nutr. Res. 2004, 24,
261–270. [CrossRef]
51. Wagenmakers, A.J. Muscle amino acid metabolism at rest and during exercise: Role in human physiology
and metabolism. Exerc. Sport Sci. Rev. 1998, 26, 287–314. [CrossRef] [PubMed]
52. Goldberg, A.L.; Chang, T.W. Regulation and significance of amino acid metabolism in skeletal muscle.
Fed. Proc. 1978, 37, 2301–2307. [PubMed]
173
Nutrients 2018, 10, 1564
53. Petry, E.R.; Cruzat, V.F.; Heck, T.G.; Leite, J.S.; Homem de Bittencourt, P.I.H.; Tirapegui, J. Alanyl-glutamine
and glutamine plus alanine supplements improve skeletal redox status in trained rats: Involvement of heat
shock protein pathways. Life Sci. 2014, 94, 130–136. [CrossRef] [PubMed]
54. Nieman, D.C.; Pedersen, B.K. Exercise and immune function. Recent developments. Sports Med. 1999, 27,
73–80. [CrossRef] [PubMed]
55. Anderson, P.M.; Broderius, M.A.; Fong, K.C.; Tsui, K.N.; Chew, S.F.; Ip, Y.K. Glutamine synthetase expression
in liver, muscle, stomach and intestine of bostrichthys sinensis in response to exposure to a high exogenous
ammonia concentration. J. Exp. Biol. 2002, 205, 2053–2065. [PubMed]
56. Austgen, T.R.; Chakrabarti, R.; Chen, M.K.; Souba, W.W. Adaptive regulation in skeletal muscle glutamine
metabolism in endotoxin-treated rats. J. Trauma 1992, 32, 600–607. [CrossRef] [PubMed]
57. Labow, B.I.; Souba, W.W.; Abcouwer, S.F. Glutamine synthetase expression in muscle is regulated by
transcriptional and posttranscriptional mechanisms. Am. J. Physiol. 1999, 276, E1136–E1145. [CrossRef]
[PubMed]
58. Xia, Y.; Wen, H.Y.; Young, M.E.; Guthrie, P.H.; Taegtmeyer, H.; Kellems, R.E. Mammalian target of rapamycin
and protein kinase a signaling mediate the cardiac transcriptional response to glutamine. J. Biolog. Chem.
2003, 278, 13143–13150. [CrossRef] [PubMed]
59. Galley, H.F. Oxidative stress and mitochondrial dysfunction in sepsis. Br. J. Anaesth. 2011, 107, 57–64.
[CrossRef] [PubMed]
60. Bode, B.P. Recent molecular advances in mammalian glutamine transport. J. Nutr. 2001, 131, 2475S–2486S.
[CrossRef] [PubMed]
61. Haussinger, D.; Schliess, F. Glutamine metabolism and signaling in the liver. Front. Biosci. 2007, 12, 371–391.
[CrossRef] [PubMed]
62. McGivan, J.D.; Bradford, N.M. Characteristics of the activation of glutaminase by ammonia in sonicated rat
liver mitochondria. Biochim. Biophys. Acta 1983, 759, 296–302. [CrossRef]
63. Hoek, J.B.; Charles, R.; De Haan, E.J.; Tager, J.M. Glutamate oxidation in rat-liver homogenate.
Biochim. Biophys. Acta 1969, 172, 407–416. [CrossRef]
64. Halestrap, A.P. The regulation of the matrix volume of mammalian mitochondria in vivo and in vitro and its
role in the control of mitochondrial metabolism. Biochim. Biophys. Acta 1989, 973, 355–382. [CrossRef]
65. Brosnan, J.T.; Brosnan, M.E. Hepatic glutaminase—A special role in urea synthesis? Nutrition 2002, 18,
455–457. [CrossRef]
66. Meijer, A.J.; Verhoeven, A.J. Regulation of hepatic glutamine metabolism. Biochem. Soc. Trans. 1986, 14,
1001–1004. [CrossRef] [PubMed]
67. Watford, M.; Smith, E.M. Distribution of hepatic glutaminase activity and mRNA in perivenous and
periportal rat hepatocytes. Biochem. J. 1990, 267, 265–267. [CrossRef] [PubMed]
68. Moorman, A.F.; de Boer, P.A.; Watford, M.; Dingemanse, M.A.; Lamers, W.H. Hepatic glutaminase mRNA
is confined to part of the urea cycle domain in the adult rodent liver lobule. FEBS Lett. 1994, 356, 76–80.
[CrossRef]
69. Gebhardt, R.; Mecke, D. Heterogeneous distribution of glutamine synthetase among rat liver parenchymal
cells in situ and in primary culture. EMBO J. 1983, 2, 567–570. [CrossRef] [PubMed]
70. Häussinger, D.; Soboll, S.; Meijer, A.J.; Gerok, W.; Tager, J.M.; Sies, H. Role of plasma membrane transport in
hepatic glutamine metabolism. Eur. J. Biochem. 1985, 152, 597–603. [CrossRef] [PubMed]
71. Lenzen, C.; Soboll, S.; Sies, H.; Haussinger, D. Ph control of hepatic glutamine degradation. Role of transport.
Eur. J. Biochem. 1987, 166, 483–488. [CrossRef] [PubMed]
72. Häussinger, D.; Hallbrucker, C.; Saha, N.; Lang, F.; Gerok, W. Cell volume and bile acid excretion. Biochem. J.
1992, 288, 681–689. [CrossRef] [PubMed]
73. Haussinger, D.; Lang, F. Cell volume in the regulation of hepatic function: A mechanism for metabolic
control. Biochim. Biophys. Acta 1991, 1071, 331–350. [CrossRef]
74. Gustafson, L.A.; Jumelle-Laclau, M.N.; van Woerkom, G.M.; van Kuilenburg, A.B.P.; Meijer, A.J. Cell
swelling and glycogen metabolism in hepatocytes from fasted rats. Biochim. Biophys. Acta 1997, 1318,
184–190. [CrossRef]
75. Baquet, A.; Gaussin, V.; Bollen, M.; Stalmans, W.; Hue, L. Mechanism of activation of liver acetyl-coa
carboxylase by cell swelling. Eur. J. Biochem. 1993, 217, 1083–1089. [CrossRef] [PubMed]
174
Nutrients 2018, 10, 1564
76. Vom Dahl, S.; Dombrowski, F.; Schmitt, M.; Schliess, F.; Pfeifer, U.; Häussinger, D. Cell hydration controls
autophagosome formation in rat liver in a microtubule-dependent way downstream from p38mapk
activation. Biochem. J. 2001, 354, 31–36. [CrossRef] [PubMed]
77. Vom Dahl, S.; Haussinger, D. Nutritional state and the swelling-induced inhibition of proteolysis in perfused
rat liver. J. Nutr. 1996, 126, 395–402. [CrossRef] [PubMed]
78. Häussinger, D.; Kubitz, R.; Reinehr, R.; Bode, J.G.; Schliess, F. Molecular aspects of medicine:
From experimental to clinical hepatology. Mol. Asp. Med. 2004, 25, 221–360. [CrossRef] [PubMed]
79. Jansen, L.T.; Adams, J.; Johnson, E.C.; Kavouras, S.A. Effects of cellular dehydration on glucose regulation in
healthy males—A pilot study. FASEB J. 2017, 31, 1014-2.
80. Friedman, S.L. Molecular regulation of hepatic fibrosis, an integrated cellular response to tissue injury. J. Biol.
Chem. 2000, 275, 2247–2250. [CrossRef] [PubMed]
81. Ghazwani, M.; Zhang, Y.; Gao, X.; Fan, J.; Li, J.; Li, S. Anti-fibrotic effect of thymoquinone on hepatic stellate
cells. Phytomedicine 2014, 21, 254–260. [CrossRef] [PubMed]
82. Li, J.; Ghazwani, M.; Liu, K.; Huang, Y.; Chang, N.; Fan, J.; He, F.; Li, L.; Bu, S.; Xie, W.; et al. Regulation of
hepatic stellate cell proliferation and activation by glutamine metabolism. PLoS ONE 2017, 12, e0182679.
[CrossRef] [PubMed]
83. Lin, Z.; Cai, F.; Lin, N.; Ye, J.; Zheng, Q.; Ding, G. Effects of glutamine on oxidative stress and nuclear
factor-κb expression in the livers of rats with nonalcoholic fatty liver disease. Exp. Ther. Med. 2014, 7,
365–370. [CrossRef] [PubMed]
84. Sellmann, C.; Baumann, A.; Brandt, A.; Jin, C.J.; Nier, A.; Bergheim, I. Oral supplementation of glutamine
attenuates the progression of nonalcoholic steatohepatitis in c57bl/6j mice. J. Nutr. 2017, 147, 2041–2049.
[CrossRef] [PubMed]
85. Magalhaes, C.R.; Malafaia, O.; Torres, O.J.; Moreira, L.B.; Tefil, S.C.; Pinherio Mda, R.; Harada, B.A. Liver
regeneration with l-glutamine supplemented diet: Experimental study in rats. Rev. Col. Bras. Cir. 2014, 41,
117–121. [CrossRef] [PubMed]
86. Helling, G.; Wahlin, S.; Smedberg, M.; Pettersson, L.; Tjäder, I.; Norberg, Å.; Rooyackers, O.; Wernerman, J.
Plasma glutamine concentrations in liver failure. PLoS ONE 2016, 11, e0150440. [CrossRef] [PubMed]
87. Eagle, H.; Oyama, V.I.; Levy, M.; Horton, C.L.; Fleischman, R. Growth response of mammalian cells in tissue
culture to l-glutamine and l-glutamic acid. J. Biol. Chem. 1956, 218, 607–616. [PubMed]
88. Newsholme, P.; Curi, R.; Gordon, S.; Newsholme, E.A. Metabolism of glucose, glutamine, long-chain
fatty-acids and ketone-bodies by murine macrophages. Biochem. J. 1986, 239, 121–125. [CrossRef] [PubMed]
89. Newsholme, E.A.; Newsholme, P.; Curi, R. The role of the citric acid cycle in cells of the immune system and
its importance in sepsis, trauma and burns. Biochem. Soc. Symp. 1987, 54, 145–162. [PubMed]
90. Curi, R.; Newsholme, P.; Newsholme, E.A. Intracellular-distribution of some enzymes of the glutamine
utilization pathway in rat lymphocytes. Biochem. Biophys. Res. Commun. 1986, 138, 318–322. [CrossRef]
91. Curi, T.C.P.; de Melo, M.P.; de Azevedo, R.B.; Curi, R. Glutamine utilisation by rat neutrophils. Biochem. Soc.
Trans. 1997, 25, 249S. [CrossRef] [PubMed]
92. Curi, T.C.P.; DeMelo, M.P.; DeAzevedo, R.B.; Zorn, T.M.T.; Curi, R. Glutamine utilization by rat neutrophils:
Presence of phosphate-dependent glutaminase. Am. J. Physiol. Cell Physiol. 1997, 273, C1124–C1129.
[CrossRef]
93. Oudemans-van Straaten, H.M.; Bosman, R.J.; Treskes, M.; van der Spoel, H.J.; Zandstra, D.F. Plasma
glutamine depletion and patient outcome in acute icu admissions. Intensiv. Care Med. 2001, 27, 84–90.
[CrossRef]
94. Leite, J.S.M.; Cruzat, V.F.; Krause, M.; Homem de Bittencourt, P.I. Physiological regulation of the heat shock
response by glutamine: Implications for chronic low-grade inflammatory diseases in age-related conditions.
Nutrire 2016, 41, 17. [CrossRef]
95. Roth, E.; Oehler, R.; Manhart, N.; Exner, R.; Wessner, B.; Strasser, E.; Spittler, A. Regulative potential of
glutamine—Relation to glutathione metabolism. Nutrition 2002, 18, 217–221. [CrossRef]
96. Hiscock, N.; Petersen, E.W.; Krzywkowski, K.; Boza, J.; Halkjaer-Kristensen, J.; Pedersen, B.K. Glutamine
supplementation further enhances exercise-induced plasma il-6. J. Appl. Physiol. 2003, 95, 145–148. [CrossRef]
[PubMed]
97. Mills, E.L.; Kelly, B.; O’Neill, L.A.J. Mitochondria are the powerhouses of immunity. Nat. Immunol. 2017, 18,
488–498. [CrossRef] [PubMed]
175
Nutrients 2018, 10, 1564
98. Pithon-Curi, T.C.; De Melo, M.P.; Curi, R. Glucose and glutamine utilization by rat lymphocytes, monocytes
and neutrophils in culture: A comparative study. Cell Biochem. Funct. 2004, 22, 321–326. [CrossRef] [PubMed]
99. Pithon-Curi, T.C.; Trezena, A.G.; Tavares-Lima, W.; Curi, R. Evidence that glutamine is involved in neutrophil
function. Cell Biochem. Funct. 2002, 20, 81–86. [CrossRef] [PubMed]
100. Branzk, N.; Lubojemska, A.; Hardison, S.E.; Wang, Q.; Gutierrez, M.G.; Brown, G.D.; Papayannopoulos, V.
Neutrophils sense microbe size and selectively release neutrophil extracellular traps in response to large
pathogens. Nat. Immunol. 2014, 15, 1017–1025. [CrossRef] [PubMed]
101. Pithon-Curi, T.C.; Levada, A.C.; Lopes, L.R.; Doi, S.Q.; Curi, R. Glutamine plays a role in superoxide
production and the expression of p47(phox), p22(phox) and gp91(phox) in rat neutrophils. Clin. Sci. 2002,
103, 403–408. [CrossRef] [PubMed]
102. Garcia, C.; Pithon-Curi, T.C.; de Lourdes Firmano, M.; Pires de Melo, M.; Newsholme, P.; Curi, R. Effects of
adrenaline on glucose and glutamine metabolism and superoxide production by rat neutrophils. Clin. Sci.
1999, 96, 549–555. [CrossRef] [PubMed]
103. Newsholme, P.; Costa Rosa, L.F.; Newsholme, E.A.; Curi, R. The importance of fuel metabolism to
macrophage function. Cell Biochem. Funct. 1996, 14, 1–10. [CrossRef] [PubMed]
104. Peres, C.M.; Procopio, J.; Costa, M.; Curi, R. Thioglycolate-elicited rat macrophages exhibit alterations in
incorporation and oxidation of fatty acids. Lipids 1999, 34, 1193–1197. [CrossRef] [PubMed]
105. Costa Rosa, L.F.; Safi, D.A.; Curi, R. Effect of thioglycollate and bcg stimuli on glucose and glutamine
metabolism in rat macrophages. J. Leukoc. Biol. 1994, 56, 10–14. [PubMed]
106. Nagy, C.; Haschemi, A. Time and demand are two critical dimensions of immunometabolism: The process
of macrophage activation and the pentose phosphate pathway. Front. Immunol. 2015, 6, 164. [CrossRef]
[PubMed]
107. Langston, P.K.; Shibata, M.; Horng, T. Metabolism supports macrophage activation. Front. Immunol. 2017, 8,
61. [CrossRef] [PubMed]
108. Vergadi, E.; Ieronymaki, E.; Lyroni, K.; Vaporidi, K.; Tsatsanis, C. Akt signaling pathway in macrophage
activation and m1/m2 polarization. J. Immunol. 2017, 198, 1006–1014. [CrossRef] [PubMed]
109. Martinez, F.O.; Sica, A.; Mantovani, A.; Locati, M. Macrophage activation and polarization. Front. Biosci.
2008, 13, 453–461. [CrossRef] [PubMed]
110. Gordon, S.; Taylor, P.R. Monocyte and macrophage heterogeneity. Nat. Rev. Immunol. 2005, 5, 953–964.
[CrossRef] [PubMed]
111. Gordon, S.; Martinez, F.O. Alternative activation of macrophages: Mechanism and functions. Immunity 2010,
32, 593–604. [CrossRef] [PubMed]
112. Mosser, D.M.; Edwards, J.P. Exploring the full spectrum of macrophage activation. Nat. Rev. Immunol. 2008,
8, 958–969. [CrossRef] [PubMed]
113. O’Neill, L.A.; Pearce, E.J. Immunometabolism governs dendritic cell and macrophage function. J. Exp. Med.
2016, 213, 15–23. [CrossRef] [PubMed]
114. Namgaladze, D.; Brune, B. Fatty acid oxidation is dispensable for human macrophage il-4-induced
polarization. Biochim. Biophys. Acta 2014, 1841, 1329–1335. [CrossRef] [PubMed]
115. O’Neill, L.A. A broken krebs cycle in macrophages. Immunity 2015, 42, 393–394. [CrossRef] [PubMed]
116. Warburg, O.; Wind, F.; Negelein, E. The metabolism of tumors in the body. J. Gen. Physiol. 1927, 8, 519–530.
[CrossRef] [PubMed]
117. Palsson-McDermott, E.M.; Curtis, A.M.; Goel, G.; Lauterbach, M.A.; Sheedy, F.J.; Gleeson, L.E.;
van den Bosch, M.W.; Quinn, S.R.; Domingo-Fernandez, R.; Johnston, D.G.; et al. Pyruvate kinase M2
regulates hif-1alpha activity and il-1beta induction and is a critical determinant of the warburg effect in
lps-activated macrophages. Cell Metab. 2015, 21, 65–80. [CrossRef] [PubMed]
118. Oren, R.; Farnham, A.E.; Saito, K.; Milofsky, E.; Karnovsky, M.L. Metabolic patterns in three types of
phagocytizing cells. J. Cell Biol. 1963, 17, 487–501. [CrossRef] [PubMed]
119. Tannahill, G.M.; Curtis, A.M.; Adamik, J.; Palsson-McDermott, E.M.; McGettrick, A.F.; Goel, G.; Frezza, C.;
Bernard, N.J.; Kelly, B.; Foley, N.H.; et al. Succinate is an inflammatory signal that induces il-1β through
hif-1α. Nature 2013, 496, 238–242. [CrossRef] [PubMed]
176
Nutrients 2018, 10, 1564
120. Jha, A.K.; Huang, S.C.; Sergushichev, A.; Lampropoulou, V.; Ivanova, Y.; Loginicheva, E.; Chmielewski, K.;
Stewart, K.M.; Ashall, J.; Everts, B.; et al. Network integration of parallel metabolic and transcriptional data
reveals metabolic modules that regulate macrophage polarization. Immunity 2015, 42, 419–430. [CrossRef]
[PubMed]
121. Davies, L.C.; Rice, C.M.; Palmieri, E.M.; Taylor, P.R.; Kuhns, D.B.; McVicar, D.W. Peritoneal tissue-resident
macrophages are metabolically poised to engage microbes using tissue-niche fuels. Nat. Commun. 2017, 8,
2074. [CrossRef] [PubMed]
122. Liu, P.S.; Wang, H.; Li, X.; Chao, T.; Teav, T.; Christen, S.; Di Conza, G.; Cheng, W.C.; Chou, C.H.;
Vavakova, M.; et al. Alpha-ketoglutarate orchestrates macrophage activation through metabolic and
epigenetic reprogramming. Nat. Immunol. 2017, 18, 985–994. [PubMed]
123. Nelson, V.L.; Nguyen, H.C.B.; Garcia-Canaveras, J.C.; Briggs, E.R.; Ho, W.Y.; DiSpirito, J.R.; Marinis, J.M.;
Hill, D.A.; Lazar, M.A. Ppargamma is a nexus controlling alternative activation of macrophages via glutamine
metabolism. Gen. Dev. 2018, 32, 1035–1044. [CrossRef] [PubMed]
124. Greiner, E.F.; Guppy, M.; Brand, K. Glucose is essential for proliferation and the glycolytic enzyme induction
that provokes a transition to glycolytic energy production. J. Biol. Chem. 1994, 269, 31484–31490. [PubMed]
125. Newsholme, E.A.; Crabtree, B.; Ardawi, M.S. Glutamine metabolism in lymphocytes: Its biochemical,
physiological and clinical importance. Q. J. Exp. Physiol. 1985, 70, 473–489. [CrossRef] [PubMed]
126. Curi, R.; Newsholme, P.; Newsholme, E.A. Metabolism of pyruvate by isolated rat mesenteric lymphocytes,
lymphocyte mitochondria and isolated mouse macrophages. Biochem. J. 1988, 250, 383–388. [CrossRef]
[PubMed]
127. Maciolek, J.A.; Pasternak, J.A.; Wilson, H.L. Metabolism of activated t lymphocytes. Curr. Opin. Immunol.
2014, 27, 60–74. [CrossRef] [PubMed]
128. Tripmacher, R.; Gaber, T.; Dziurla, R.; Haupl, T.; Erekul, K.; Grutzkau, A.; Tschirschmann, M.; Scheffold, A.;
Radbruch, A.; Burmester, G.R.; et al. Human cd4(+) T cells maintain specific functions even under conditions
of extremely restricted ATP production. Eur. J. Immunol. 2008, 38, 1631–1642. [CrossRef] [PubMed]
129. Wieman, H.L.; Wofford, J.A.; Rathmell, J.C. Cytokine stimulation promotes glucose uptake via
phosphatidylinositol-3 kinase/akt regulation of glut1 activity and trafficking. Mol. Biol. Cell 2007, 18,
1437–1446. [CrossRef] [PubMed]
130. Delgoffe, G.M.; Kole, T.P.; Zheng, Y.; Zarek, P.E.; Matthews, K.L.; Xiao, B.; Worley, P.F.; Kozma, S.C.;
Powell, J.D. The mtor kinase differentially regulates effector and regulatory T cell lineage commitment.
Immunity 2009, 30, 832–844. [CrossRef] [PubMed]
131. Lee, K.; Gudapati, P.; Dragovic, S.; Spencer, C.; Joyce, S.; Killeen, N.; Magnuson, M.A.; Boothby, M.
Mammalian target of rapamycin protein complex 2 regulates differentiation of th1 and th2 cell subsets
via distinct signaling pathways. Immunity 2010, 32, 743–753. [CrossRef] [PubMed]
132. Michalek, R.D.; Gerriets, V.A.; Jacobs, S.R.; Macintyre, A.N.; MacIver, N.J.; Mason, E.F.; Sullivan, S.A.;
Nichols, A.G.; Rathmell, J.C. Cutting edge: Distinct glycolytic and lipid oxidative metabolic programs
are essential for effector and regulatory cd4+ T cell subsets. J. Immunol. 2011, 186, 3299–3303. [CrossRef]
[PubMed]
133. Hardie, D.G.; Hawley, S.A.; Scott, J.W. Amp-activated protein kinas—Development of the energy sensor
concept. J. Physiol. 2006, 574, 7–15. [CrossRef] [PubMed]
134. Crawford, J.; Cohen, H.J. The essential role of l-glutamine in lymphocyte differentiation in vitro. J. Cell.
Physiol. 1985, 124, 275–282. [CrossRef] [PubMed]
135. Matarese, G.; Colamatteo, A.; De Rosa, V. Metabolic fuelling of proper t cell functions. Immunol. Lett. 2014,
161, 174–178. [CrossRef] [PubMed]
136. Chang, C.H.; Curtis, J.D.; Maggi, L.B., Jr.; Faubert, B.; Villarino, A.V.; O’Sullivan, D.; Huang, S.C.;
van der Windt, G.J.; Blagih, J.; Qiu, J.; et al. Posttranscriptional control of t cell effector function by aerobic
glycolysis. Cell 2013, 153, 1239–1251. [CrossRef] [PubMed]
137. Zheng, Y.; Delgoffe, G.M.; Meyer, C.F.; Chan, W.; Powell, J.D. Anergic T cells are metabolically anergic.
J. Immunol. 2009, 183, 6095–6101. [CrossRef] [PubMed]
138. Buck, M.D.; O’Sullivan, D.; Klein Geltink, R.I.; Curtis, J.D.; Chang, C.H.; Sanin, D.E.; Qiu, J.; Kretz, O.;
Braas, D.; van der Windt, G.J.; et al. Mitochondrial dynamics controls T cell fate through metabolic
programming. Cell 2016, 166, 63–76. [CrossRef] [PubMed]
177
Nutrients 2018, 10, 1564
139. Corcoran, S.E.; O’Neill, L.A. Hif1alpha and metabolic reprogramming in inflammation. J. Clin. Investig. 2016,
126, 3699–3707. [CrossRef] [PubMed]
140. Araujo, L.; Khim, P.; Mkhikian, H.; Mortales, C.L.; Demetriou, M. Glycolysis and glutaminolysis
cooperatively control T cell function by limiting metabolite supply to n-glycosylation. eLife 2017, 6, e21330.
[CrossRef] [PubMed]
141. Hesterberg, R.S.; Cleveland, J.L.; Epling-Burnette, P.K. Role of polyamines in immune cell functions. Med. Sci.
2018, 6, 22. [CrossRef] [PubMed]
142. Calder, P.C.; Yaqoob, P. Glutamine and the immune system. Amino Acids 1999, 17, 227–241. [CrossRef]
[PubMed]
143. Wilmore, D.W.; Shabert, J.K. Role of glutamine in immunologic responses. Nutrition 1998, 14, 618–626.
[CrossRef]
144. Lagranha, C.J.; Hirabara, S.M.; Curi, R.; Pithon-Curi, T.C. Glutamine supplementation prevents
exercise-induced neutrophil apoptosis and reduces p38 mapk and jnk phosphorylation and p53 and caspase
3 expression. Cell Biochem. Funct. 2007, 25, 563–569. [CrossRef] [PubMed]
145. Young, V.R.; Ajami, A.M. Glutamine: The emperor or his clothes? J. Nutr. 2001, 131, 2447S–2486S. [CrossRef]
[PubMed]
146. Meister, A.; Anderson, M.E. Glutathione. Ann. Rev. Biochem. 1983, 52, 711–760. [CrossRef] [PubMed]
147. Gaucher, C.; Boudier, A.; Bonetti, J.; Clarot, I.; Leroy, P.; Parent, M. Glutathione: Antioxidant properties
dedicated to nanotechnologies. Antioxidants 2018, 7, 62. [CrossRef] [PubMed]
148. Liu, N.; Ma, X.; Luo, X.; Zhang, Y.; He, Y.; Dai, Z.; Yang, Y.; Wu, G.; Wu, Z. L-glutamine attenuates apoptosis
in porcine enterocytes by regulating glutathione-related redox homeostasis. J. Nutr. 2018, 148, 526–534.
[CrossRef] [PubMed]
149. Da Silva Lima, F.; Rogero, M.M.; Ramos, M.C.; Borelli, P.; Fock, R.A. Modulation of the nuclear factor-kappa
b (nf-kappab) signalling pathway by glutamine in peritoneal macrophages of a murine model of protein
malnutrition. Eur. J. Nutr. 2013, 52, 1343–1351. [CrossRef] [PubMed]
150. Smedberg, M.; Wernerman, J. Is the glutamine story over? Crit. Care 2016, 20, 361. [CrossRef] [PubMed]
151. Heck, T.G.; Scholer, C.M.; de Bittencourt, P.I. Hsp70 expression: Does it a novel fatigue signalling factor from
immune system to the brain? Cell Biochem. Funct. 2011, 29, 215–226. [CrossRef] [PubMed]
152. Singleton, K.D.; Wischmeyer, P.E. Glutamine’s protection against sepsis and lung injury is dependent on heat
shock protein 70 expression. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2007, 292, R1839–R1845. [CrossRef]
[PubMed]
153. Jordan, I.; Balaguer, M.; Esteban, M.E.; Cambra, F.J.; Felipe, A.; Hernandez, L.; Alsina, L.; Molero, M.;
Villaronga, M.; Esteban, E. Glutamine effects on heat shock protein 70 and interleukines 6 and 10:
Randomized trial of glutamine supplementation versus standard parenteral nutrition in critically ill children.
Clin. Nutr. 2016, 35, 34–40. [CrossRef] [PubMed]
154. Kim, G.; Meriin, A.B.; Gabai, V.L.; Christians, E.; Benjamin, I.; Wilson, A.; Wolozin, B.; Sherman, M.Y. The heat
shock transcription factor hsf1 is downregulated in DNA damage-associated senescence, contributing to the
maintenance of senescence phenotype. Aging Cell 2012, 11, 617–627. [CrossRef] [PubMed]
155. Gabai, V.L.; Meng, L.; Kim, G.; Mills, T.A.; Benjamin, I.J.; Sherman, M.Y. Heat shock transcription factor hsf1
is involved in tumor progression via regulation of hypoxia-inducible factor 1 and RNA-binding protein
HUR. Mol. Cell. Biol. 2012, 32, 929–940. [CrossRef] [PubMed]
156. Dokladny, K.; Zuhl, M.N.; Mandell, M.; Bhattacharya, D.; Schneider, S.; Deretic, V.; Moseley, P.L. Regulatory
coordination between two major intracellular homeostatic systems: Heat shock response and autophagy.
J. Biolog. Chem. 2013, 288, 14959–14972. [CrossRef] [PubMed]
157. Martinez, M.R.; Dias, T.B.; Natov, P.S.; Zachara, N.E. Stress-induced o-glcnacylation: An adaptive process of
injured cells. Biochem. Soc. Trans. 2017, 45, 237–249. [CrossRef] [PubMed]
158. Lafontaine-Lacasse, M.; Dore, G.; Picard, F. Hexosamines stimulate apoptosis by altering sirt1 action and
levels in rodent pancreatic beta-cells. J. Endoc. 2011, 208, 41–49. [CrossRef] [PubMed]
159. Kazemi, Z.; Chang, H.; Haserodt, S.; McKen, C.; Zachara, N.E. O-linked beta-n-acetylglucosamine (o-glcnac)
regulates stress-induced heat shock protein expression in a gsk-3beta-dependent manner. J. Biol. Chem. 2010,
285, 39096–39107. [CrossRef] [PubMed]
178
Nutrients 2018, 10, 1564
160. Hamiel, C.R.; Pinto, S.; Hau, A.; Wischmeyer, P.E. Glutamine enhances heat shock protein 70 expression
via increased hexosamine biosynthetic pathway activity. Am. J. Physiol. Cell Physiol. 2009, 297, C1509–1519.
[CrossRef] [PubMed]
161. Singleton, K.D.; Serkova, N.; Beckey, V.E.; Wischmeyer, P.E. Glutamine attenuates lung injury and improves
survival after sepsis: Role of enhanced heat shock protein expression. Crit. Care Med. 2005, 33, 1206–1213.
[CrossRef] [PubMed]
162. Raizel, R.; Leite, J.S.; Hypolito, T.M.; Coqueiro, A.Y.; Newsholme, P.; Cruzat, V.F.; Tirapegui, J.
Determination of the anti-inflammatory and cytoprotective effects of l-glutamine and l-alanine, or dipeptide,
supplementation in rats submitted to resistance exercise. Br. J. Nutr. 2016, 116, 470–479. [CrossRef] [PubMed]
163. Smolka, M.B.; Zoppi, C.C.; Alves, A.A.; Silveira, L.R.; Marangoni, S.; Pereira-Da-Silva, L.; Novello, J.C.;
Macedo, D.V. Hsp72 as a complementary protection against oxidative stress induced by exercise in the soleus
muscle of rats. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2000, 279, R1539–R1545. [CrossRef] [PubMed]
164. Gupta, A.; Cooper, Z.A.; Tulapurkar, M.E.; Potla, R.; Maity, T.; Hasday, J.D.; Singh, I.S. Toll-like receptor
agonists and febrile range hyperthermia synergize to induce heat shock protein 70 expression and
extracellular release. J. Biolog. Chem. 2013, 288, 2756–2766. [CrossRef] [PubMed]
165. Krause, M.; Keane, K.; Rodrigues-Krause, J.; Crognale, D.; Egan, B.; De Vito, G.; Murphy, C.; Newsholme, P.
Elevated levels of extracellular heat-shock protein 72 (ehsp72) are positively correlated with insulin resistance
in vivo and cause pancreatic beta-cell dysfunction and death in vitro. Clin. Sci. 2014, 126, 739–752. [CrossRef]
[PubMed]
166. Lenders, C.M.; Liu, S.; Wilmore, D.W.; Sampson, L.; Dougherty, L.W.; Spiegelman, D.; Willett, W.C. Evaluation
of a novel food composition database that includes glutamine and other amino acids derived from gene
sequencing data. Eur. J. Clin. Nutr. 2009, 63, 1433–1439. [CrossRef] [PubMed]
167. Hermans, G.; Van den Berghe, G. Clinical review: Intensive care unit acquired weakness. Crit. Care 2015, 19,
274. [CrossRef] [PubMed]
168. Stehle, P.; Ellger, B.; Kojic, D.; Feuersenger, A.; Schneid, C.; Stover, J.; Scheiner, D.; Westphal, M. Glutamine
dipeptide-supplemented parenteral nutrition improves the clinical outcomes of critically ill patients:
A systematic evaluation of randomised controlled trials. Clin. Nutr. ESPEN 2017, 17, 75–85. [CrossRef]
[PubMed]
169. Gunst, J.; Vanhorebeek, I.; Thiessen, S.E.; Van den Berghe, G. Amino acid supplements in critically ill patients.
Pharmacol. Res. 2018, 130, 127–131. [CrossRef] [PubMed]
170. Furst, P.; Alteheld, B.; Stehle, P. Why should a single nutrient—Glutamine—Improve outcome?
The remarkable story of glutamine dipeptides. Clin. Nutr. Suppl. 2004, 1, 3–15. [CrossRef]
171. Grau, T.; Bonet, A.; Minambres, E.; Pineiro, L.; Irles, J.A.; Robles, A.; Acosta, J.; Herrero, I.; Palacios, V.;
Lopez, J.; et al. The effect of l-alanyl-l-glutamine dipeptide supplemented total parenteral nutrition on
infectious morbidity and insulin sensitivity in critically ill patients. Crit. Care Med. 2011, 39, 1263–1268.
[CrossRef] [PubMed]
172. Estivariz, C.F.; Griffith, D.P.; Luo, M.; Szeszycki, E.E.; Bazargan, N.; Dave, N.; Daignault, N.M.; Bergman, G.F.;
McNally, T.; Battey, C.H.; et al. Efficacy of parenteral nutrition supplemented with glutamine dipeptide
to decrease hospital infections in critically ill surgical patients. J. Parenter. Enter. Nutr. 2008, 32, 389–402.
[CrossRef] [PubMed]
173. Wang, Y.; Jiang, Z.M.; Nolan, M.T.; Jiang, H.; Han, H.R.; Yu, K.; Li, H.L.; Jie, B.; Liang, X.K. The impact of
glutamine dipeptide-supplemented parenteral nutrition on outcomes of surgical patients: A meta-analysis
of randomized clinical trials. J. Parenter. Enter. Nutr. 2010, 34, 521–529. [CrossRef] [PubMed]
174. Bollhalder, L.; Pfeil, A.M.; Tomonaga, Y.; Schwenkglenks, M. A systematic literature review and meta-analysis
of randomized clinical trials of parenteral glutamine supplementation. Clin. Nutr. 2013, 32, 213–223.
[CrossRef] [PubMed]
175. Dechelotte, P.; Hasselmann, M.; Cynober, L.; Allaouchiche, B.; Coeffier, M.; Hecketsweiler, B.; Merle, V.;
Mazerolles, M.; Samba, D.; Guillou, Y.M.; et al. L-alanyl-L-glutamine dipeptide-supplemented total parenteral
nutrition reduces infectious complications and glucose intolerance in critically ill patients: The french
controlled, randomized, double-blind, multicenter study. Crit. Care Med. 2006, 34, 598–604. [CrossRef]
[PubMed]
176. Weitzel, L.R.; Wischmeyer, P.E. Glutamine in critical illness: The time has come, the time is now. Crit. Care
Clin. 2010, 26, 515–525. [CrossRef] [PubMed]
179
Nutrients 2018, 10, 1564
177. Klassen, P.; Mazariegos, M.; Solomons, N.W.; Furst, P. The pharmacokinetic responses of humans to 20 g of
alanyl-glutamine dipeptide differ with the dosing protocol but not with gastric acidity or in patients with
acute dengue fever. J. Nutr. 2000, 130, 177–182. [CrossRef] [PubMed]
178. Melis, G.C.; Boelens, P.G.; van der Sijp, J.R.; Popovici, T.; De Bandt, J.P.; Cynober, L.; van Leeuwen, P.A.
The feeding route (enteral or parenteral) affects the plasma response of the dipetide ala-gln and the amino
acids glutamine, citrulline and arginine, with the administration of ala-gln in preoperative patients. Br. J. Nutr.
2005, 94, 19–26. [CrossRef] [PubMed]
179. Krause, M.S.; de Bittencourt, P.I.H.J. Type 1 diabetes: Can exercise impair the autoimmune event?
The l-arginine/glutamine coupling hypothesis. Cell Biochem. Funct. 2008, 26, 406–433. [CrossRef] [PubMed]
180. Adibi, S.A. Regulation of expression of the intestinal oligopeptide transporter (pept-1) in health and disease.
Am. J. Physiol. Gastrointest. Liver Physiol. 2003, 285, G779–G788. [CrossRef] [PubMed]
181. Broer, S. Amino acid transport across mammalian intestinal and renal epithelia. Physiol. Rev. 2008, 88,
249–286. [CrossRef] [PubMed]
182. Gilbert, E.R.; Wong, E.A.; Webb, K.E. Board-invited review: Peptide absorption and utilization: Implications
for animal nutrition and health. J. Anim. Sci. 2008, 86, 2135–2155. [CrossRef] [PubMed]
183. Petry, E.R.; Cruzat, V.F.; Heck, T.G.; de Bittencourt, P.I.H.; Tirapegui, J. L-glutamine supplementations
enhance liver glutamine-glutathione axis and heat shock factor-1 expression in endurance-exercise trained
rats. Int. J. Sport Nutr. Exerc. Metab. 2015, 25, 188–197. [CrossRef] [PubMed]
184. Alba-Loureiro, T.C.; Ribeiro, R.F.; Zorn, T.M.; Lagranha, C.J. Effects of glutamine supplementation on kidney
of diabetic rat. Amino Acids 2010, 38, 1021–1030. [CrossRef] [PubMed]
185. Cheng, T.; Sudderth, J.; Yang, C.; Mullen, A.R.; Jin, E.S.; Mates, J.M.; DeBerardinis, R.J. Pyruvate carboxylase
is required for glutamine-independent growth of tumor cells. Proc. Natl. Acad Sci. USA 2011, 108, 8674–8679.
[CrossRef] [PubMed]
186. Hensley, C.T.; Wasti, A.T.; DeBerardinis, R.J. Glutamine and cancer: Cell biology, physiology, and clinical
opportunities. J. Clin. Investig. 2013, 123, 3678–3684. [CrossRef] [PubMed]
187. Hensley, C.T.; Faubert, B.; Yuan, Q.; Lev-Cohain, N.; Jin, E.; Kim, J.; Jiang, L.; Ko, B.; Skelton, R.; Loudat, L.;
et al. Metabolic heterogeneity in human lung tumors. Cell 2016, 164, 681–694. [CrossRef] [PubMed]
188. Davidson, S.M.; Papagiannakopoulos, T.; Olenchock, B.A.; Heyman, J.E.; Keibler, M.A.; Luengo, A.;
Bauer, M.R.; Jha, A.K.; O’Brien, J.P.; Pierce, K.A.; et al. Environment impacts the metabolic dependencies of
ras-driven non-small cell lung cancer. Cell Metab. 2016, 23, 517–528. [CrossRef] [PubMed]
189. Choi, C.; Ganji, S.; Hulsey, K.; Madan, A.; Kovacs, Z.; Dimitrov, I.; Zhang, S.; Pichumani, K.; Mendelsohn, D.;
Mickey, B.; et al. A comparative study of short- and long-te (1) h mrs at 3 t for in vivo detection of
2-hydroxyglutarate in brain tumors. NMR Biomed. 2013, 26, 1242–1250. [CrossRef] [PubMed]
190. Tardito, S.; Oudin, A.; Ahmed, S.U.; Fack, F.; Keunen, O.; Zheng, L.; Miletic, H.; Sakariassen, P.O.;
Weinstock, A.; Wagner, A.; et al. Glutamine synthetase activity fuels nucleotide biosynthesis and supports
growth of glutamine-restricted glioblastoma. Nat. Cell Biol. 2015, 17, 1556–1568. [CrossRef] [PubMed]
191. Deep, G.; Schlaepfer, I.R. Aberrant lipid metabolism promotes prostate cancer: Role in cell survival under
hypoxia and extracellular vesicles biogenesis. Int. J. Mol. Sci. 2016, 17, 1061. [CrossRef] [PubMed]
192. White, M.A.; Lin, C.; Rajapakshe, K.; Dong, J.; Shi, Y.; Tsouko, E.; Mukhopadhyay, R.; Jasso, D.; Dawood, W.;
Coarfa, C.; et al. Glutamine transporters are targets of multiple oncogenic signaling pathways in prostate
cancer. Mol. Cancer Res. 2017, 15, 1017–1028. [CrossRef] [PubMed]
193. Marian, M.J. Dietary supplements commonly used by cancer survivors: Are there any benefits? Nutr. Clin.
Pract. 2017, 32, 607–627. [CrossRef] [PubMed]
194. Sayles, C.; Hickerson, S.C.; Bhat, R.R.; Hall, J.; Garey, K.W.; Trivedi, M.V. Oral glutamine in preventing
treatment-related mucositis in adult patients with cancer: A systematic review. Nutr. Clin. Pract. 2016, 31,
171–179. [CrossRef] [PubMed]
195. Daniele, B.; Perrone, F.; Gallo, C.; Pignata, S.; De Martino, S.; De Vivo, R.; Barletta, E.; Tambaro, R.; Abbiati, R.;
D’Agostino, L. Oral glutamine in the prevention of fluorouracil induced intestinal toxicity: A double blind,
placebo controlled, randomised trial. Gut 2001, 48, 28–33. [CrossRef] [PubMed]
196. Hammarqvist, F.; Wernerman, J.; Ali, R.; von der Decken, A.; Vinnars, E. Addition of glutamine to total
parenteral nutrition after elective abdominal surgery spares free glutamine in muscle, counteracts the fall
in muscle protein synthesis, and improves nitrogen balance. Ann. Surg. 1989, 209, 455–461. [CrossRef]
[PubMed]
180
Nutrients 2018, 10, 1564
197. Souba, W.W.; Herskowitz, K.; Klimberg, V.S.; Salloum, R.M.; Plumley, D.A.; Flynn, T.C.; Copeland, E.M.
The effects of sepsis and endotoxemia on gut glutamine metabolism. Ann. Surg. 1990, 211, 543–551.
[CrossRef] [PubMed]
198. Bode, B.P.; Fuchs, B.C.; Hurley, B.P.; Conroy, J.L.; Suetterlin, J.E.; Tanabe, K.K.; Rhoads, D.B.; Abcouwer, S.F.;
Souba, W.W. Molecular and functional analysis of glutamine uptake in human hepatoma and liver-derived
cells. Am. J. Physiol. Gastrointest. Liver Physiol. 2002, 283, G1062–G1073. [CrossRef] [PubMed]
199. Rogero, M.M.; Borelli, P.; Fock, R.A.; Borges, M.C.; Vinolo, M.A.R.; Curi, R.; Nakajima, K.; Crisma, A.R.;
Ramos, A.D.; Tirapegui, J. Effects of glutamine on the nuclear factor-kappab signaling pathway of murine
peritoneal macrophages. Amino Acids 2010, 39, 435–441. [CrossRef] [PubMed]
200. Parry-Billings, M.; Evans, J.; Calder, P.C.; Newsholme, E.A. Does glutamine contribute to immunosuppression
after major burns? Lancet 1990, 336, 523–525. [CrossRef]
201. Roth, E.; Funovics, J.; Muhlbacher, F.; Schemper, M.; Mauritz, W.; Sporn, P.; Fritsch, A. Metabolic disorders in
severe abdominal sepsis: Glutamine deficiency in skeletal muscle. Clin. Nutr. 1982, 1, 25–41. [CrossRef]
© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
181
nutrients
Review
Immune Function and Micronutrient Requirements
Change over the Life Course
Silvia Maggini 1, *, Adeline Pierre 1 and Philip C. Calder 2,3
1 Bayer Consumer Care AG, 4002 Basel, Switzerland; adeline.pierre@bayer.com
2 Human Development & Health, Faculty of Medicine, University of Southampton,
Southampton SO16 6YD, UK; P.C.Calder@soton.ac.uk
3 NIHR Southampton Biomedical Research Centre, University Hospital Southampton NHS Foundation Trust
and University of Southampton, Southampton SO16 6YD, UK
* Correspondence: silvia.maggini@bayer.com; Tel.: +41-582-727-516
Abstract: As humans age, the risk and severity of infections vary in line with immune competence
according to how the immune system develops, matures, and declines. Several factors influence
the immune system and its competence, including nutrition. A bidirectional relationship among
nutrition, infection and immunity exists: changes in one component affect the others. For example,
distinct immune features present during each life stage may affect the type, prevalence, and severity
of infections, while poor nutrition can compromise immune function and increase infection risk.
Various micronutrients are essential for immunocompetence, particularly vitamins A, C, D, E, B2, B6,
and B12, folic acid, iron, selenium, and zinc. Micronutrient deficiencies are a recognized global public
health issue, and poor nutritional status predisposes to certain infections. Immune function may be
improved by restoring deficient micronutrients to recommended levels, thereby increasing resistance
to infection and supporting faster recovery when infected. Diet alone may be insufficient and tailored
micronutrient supplementation based on specific age-related needs necessary. This review looks at
immune considerations specific to each life stage, the consequent risk of infection, micronutrient
requirements and deficiencies exhibited over the life course, and the available evidence regarding the
effects of micronutrient supplementation on immune function and infection.
1. Introduction
The immune system, which is integrated into all physiological systems, protects the body against
infections and other external and internal insults by utilizing three distinct layers, depending on the
nature of the threat: physical (e.g., skin, epithelial lining of the gastrointestinal and respiratory tracts)
and biochemical barriers (e.g., secretions, mucus, and gastric acid), numerous different immune cells
(e.g., granulocytes, CD4 or CD8 T and B cells), and antibodies (i.e., immunoglobulins). The first line of
defense is innate immunity, which combines physical and biochemical barriers with a non-specific,
leukocyte-mediated cellular response to defend against pathogens [1]. If the pathogen manages to
avoid these innate defenses, a more complex, adaptive, antigen-specific response is triggered, mediated
by T and B lymphocytes, which produces antibodies to target and destroy the pathogen (Figure 1) [1].
Both systems also protect against native cells that may be harmful, such as cancerous or precancerous
cells [2].
Figure 1. Simple overview of the immune system. The three layers of the immune system (physical
and biochemical barriers; cells such as monocytes, granulocytes, lymphocytes, and B and T cells; and
antibodies or immunoglobulins) work together to protect the body against pathogens, utilizing the
innate and adaptive defense mechanisms. All three layers are involved in the innate and immune
systems. * The innate immune system comprises anatomical and biochemical barriers and an unspecific
cellular response mediated mainly by monocytes, neutrophils, natural killer cells and dendritic cells;
these work together to fight off pathogens before they can start an active infection. ** The adaptive
immune system involves an antigen-specific response mediated by T and B lymphocytes that is
activated by exposure to pathogens; this works with the innate immune system to reduce the severity
of infection. a The complement system can work with both the innate and adaptive immune systems;
b i.e., immunity from serum antibodies produced by plasma cells; c i.e., an immune response that
does not involve antibodies, but responds to any cells that display aberrant major histocompatibility
complex (MHC) markers, such as cells invaded by pathogens.
As humans age, the immune system evolves from the immature and developing immune
responses in infants and children, through to immune function that is potentially optimal in adolescents
and young adults, followed by a gradual decline in immunity (particularly adaptive processes) in older
people [1]. Age-related changes are compounded by certain lifestyle factors (e.g., diet, environmental
factors, and oxidative stress) specific to each life stage that can influence and modify, in some cases
suppressing, immune function. Accordingly, the risk and severity of infections such as the common
cold and influenza (the most common illnesses in humans [3]), pneumonia and diarrheal infections
also vary over a lifetime.
Optimal immune function is dependent on a healthy immune system. In turn, adequate nutrition
is crucial to ensure a good supply of the energy sources, macronutrients and micronutrients required
for the development, maintenance and expression of the immune response [3]. Micronutrients have
vital roles throughout the immune system that are independent of life stage (Table 1), and it has been
determined that those most needed to sustain immunocompetence include vitamins A, C, D, E, B2,
B6 and B12, folic acid, beta carotene, iron, selenium, and zinc [4]. There is a bidirectional interaction
among nutrition, infection and immunity: the immune response is compromised when nutrition
is poor, predisposing individuals to infections, and a poor nutritional state may be exacerbated by
the immune response itself to an infection [5]. It is clear that optimal immunocompetence depends
upon nutritional status [6]. It is recognized that micronutrient deficiencies and suboptimal intakes are
183
Nutrients 2018, 10, 1531
common worldwide [7], and certain micronutrients may be more likely to be insufficient at different
stages of the life course. This can affect the risk and severity of infection, and in fact an individual’s
nutritional status can predict the clinical course and outcome of certain infections such as diarrhea,
pneumonia and measles [4]. Resistance to infection may be enhanced by adding the deficient nutrient
back into the diet and restoring immune function [4]. However, it is not always possible to achieve
good nutritional status via the diet alone. In developing countries, for example, it may be difficult to
find an adequate and varied supply of food. Even in industrialized nations, where it may be presumed
that healthy, nutritious food is easier to obtain, social, economic, educational, ethnic and cultural
backgrounds influence the diet and may adversely affect an individual’s micronutrient status [8].
This review looks at life-stage-specific immunity, risk of infection and micronutrient requirements,
from the perspective of industrialized countries where possible. The aim is to highlight the role of
tailored supplementation in restoring micronutrients to recommended levels and better supporting
immune needs that are specific to each life stage.
184
Table 1. Overview of key roles played by select micronutrients in the immune system [4,9–14].
185
helps protect against infection caused by pathogens [14]
cells and inhibits T cell proliferation [14]
1,25-dihydroxyvitamin D3 , the active form of vitamin D, regulates
the antimicrobial proteins cathelicidin and defensin, which can
directly kill pathogens, especially bacteria [14]
Helps maintain structural and functional integrity of mucosal cells Necessary for proper functioning of T and B lymphocytes,
in innate barriers (e.g., skin, respiratory tract, etc.) [14] and thus for generation of antibody responses to antigen [14]
Vitamin A
Important for normal functioning of innate immune cells (e.g., NK Involved in development and differentiation of Th1 and Th2 cells
cells, macrophages, neutrophils) [14] and supports Th2 anti-inflammatory response [10]
An important fat-soluble antioxidant [10]
Enhances T cell-mediated functions and lymphocyte
Protects the integrity of cell membranes from damage caused by
Vitamin E proliferation [10]
free radicals [14]
Optimizes and enhances Th1 and suppresses Th2 response [10]
Enhances IL-2 production and NK cell cytotoxic activity [10]
Required in the endogenous synthesis and metabolism of amino
acids, the building blocks of cytokines and antibodies [14]
Helps regulate inflammation [13] Has roles in lymphocyte proliferation, differentiation and
Vitamin B6
Has roles in cytokine production and NK cell activity [13,15] maturation [14]
Maintains Th1 immune response [10]
Has roles in antibody production [13]
Table 1. Cont.
186
Iron Essential for cell differentiation and growth, component of
of killing bacteria by neutrophils [10]
enzymes critical for functioning of immune cells (e.g.,
Important in the generation of ROS that kill pathogens [14]
ribonucleotide reductase involved in DNA synthesis) [10]
Free-radical scavenger [4]
Antimicrobial properties [14]
Accumulates at sites of inflammation, important for IL-2 Has roles in T cell proliferation [13]
Copper
production and response [13,14] Has roles in antibody production and cellular immunity [18]
May play a role in the innate immune response to bacterial
infections [14]
Essential for the function of selenium-dependent enzymes
(selenoproteins) that can act as redox regulators and cellular Involved in T lymphocyte proliferation [4,13]
Selenium antioxidants, potentially counteracting ROS [10,14] Has roles in the humoral system (e.g., immunoglobulin
Selenoproteins are important for the antioxidant host defense production) [13]
system affecting leukocyte and NK cell function [13]
IL, interleukin; NK, natural killer; RNS, reactive nitrogen species; ROS, reactive oxygen species; Th, helper T cell.
Nutrients 2018, 10, 1531
187
Nutrients 2018, 10, 1531
concentrations (both then decrease with age); lymphocyte and platelet counts are lower in children
compared with infants and steadily decline with age [25]. Closer analysis of lymphocyte subtypes
indicates that the proportion of different lymphocyte subsets changes over time [25]. For example,
the percentage of CD3+ T cells (required for activation of CD4+ and CD8+ T cells) is significantly higher
in children than in infants. However, the proportion of CD4+ T cells is significantly lower in children
than in infants [25]. CD4+ helper T cells recognize peptides presented by major histocompatibility
complex (MHC) II molecules found on antigen-presenting cells, and subsequently secrete cytokines
that facilitate different immune responses according to the source of the antigen [27]. In contrast,
the percentage of CD8+ T cells is significantly higher in children than in infants and steadily increases
over time [25]. CD8+ cytotoxic T cells recognize peptides presented by MHC I molecules found on all
nucleated cells, and secrete cytokines like tumor-necrosis factor alpha or interferon gamma to help to
kill infected or malignant cells [27]. Analysis of B cells indicates that the proportion of CD19+ cells is
highest in infants and children and decreases significantly thereafter [25]. CD19 is an antigen that is
present on all B cells, is involved in signaling, and is a biomarker for B lymphocyte development [28].
Antibody production increases with age from infancy to childhood. For example, adult levels of
IgG (expressed on the surface of mature B cells, and the most prevalent immunoglobulin in serum)
are reached by the age of 11–12 years, with a further increase during puberty, while levels of IgA
(the second most prevalent immunoglobulin in serum, which can activate the complement pathway)
continue to increase past puberty until they reach adult levels; in contrast, adult levels of IgM (the first
immunoglobulin made by the fetus and virgin B cells challenged with antigen) are reached by the age
of four years [29].
188
Nutrients 2018, 10, 1531
effects of estrogen in women and the humoral immunity suppressing effects of testosterone in men;
however, the full extent of sex on functional immune responses remains unclear [23].
189
Nutrients 2018, 10, 1531
3. Response to Infection
The nature of the response of the immune system to a pathogen is initially dependent on whether
the innate immune defenses can eliminate the infectious organism. If not, previous experience with
the pathogen will determine how rapidly T and B cells in the adaptive immune system are able to
mount a defense against it, supported by the innate immune system. Certain factors may affect the
response of the immune system to infection.
190
Nutrients 2018, 10, 1531
pneumonia) are more common in children under five years old than any other age group worldwide,
and risk factors include air pollution and suboptimal breastfeeding [48]. Micronutrient deficiencies
also have immunological consequences in infants and young children, and can increase morbidity and
mortality from many diseases, including pneumonia, diarrheal disease, and measles [4,49]. Infection
and undernutrition have a synergistic relationship, and micronutrient deficiencies cause specific
immune impairments that affect both the innate and adaptive immune systems, such as impaired
phagocyte and lymphocyte activity with zinc deficiency, or compromised development of neutrophils,
macrophages and NK cells with vitamin A deficiency [50].
191
Nutrients 2018, 10, 1531
Figure 2. Life-style factors affecting immune function during adulthood. The risk of infection is also
influenced by gender, early programming, vaccination history, pathogen exposure, specific health
conditions, and diseases.
192
Nutrients 2018, 10, 1531
193
Table 2. Life-stage-specific micronutrient deficiencies in Europe. Reported micronutrient intakes that are below the recommended dietary allowance are shown
in bold. The table also shows the tolerable upper intake levels, the highest level of daily nutrient intake that is likely to pose no risk of adverse health effects in
most people.
Recommended Dietary Allowance [78] Tolerable Upper Intake Levels [78] Reported Mean Micronutrient Intakes, Min–Max [96]
Children a a Children
Select Children
4–8 years Adults Older age 4–6 years: M/F Older age
Nutrients 2018, 10, 1531
194
900/700 2800/2800
400–1800/300–1600
5.3–9.8/5.1–9.8
7 300
Vitamin E, 6.3–11.2/5.9–13.3
11 15 15 600 1000 1000 3.3–17.7/4.2–16.1 6.3–13.7/6.7–13.7
mg/day 5.9–14.5/5.6–18.1
15 800
6.8–20.8/6.0–15.5
1.3–1.8/1.0–1.9
0.6 40
Vitamin B6, 1.2–2.5/1.1–1.9
1.0 1.3 1.7/1.5 60 100 100 1.6–3.5/1.3–2.1 1.2–3.0/1.2–2.9
mg/day 1.2–2.8/1.1–2.7
1.3/1.2 80
1.5–3.1/1.2–2.5
2.7–5.3/2.6–5.0
1.2
Vitamin B12, 3.6–5.5/2.2–5.3
1.8 2.4 2.4 ND ND ND 1.9–9.3/1.0–8.8 3.1–8.2/2.5–7.5
μg/day 3.2–11.8/2.2–11.1
2.4
4.9–7.5/3.5–5.2
120–256/109–199
200 400
144–290/133–264
Folate, μg/day 300 300-400 400 600 1000 1000 203–494/131–392 139–343/121–335
149–428/140–360
400 800
190–365/154–298
Table 2. Cont.
Recommended Dietary Allowance [78] Tolerable Upper Intake Levels [78] Reported Mean Micronutrient Intakes, Min–Max [96]
Children a a Children
Select Children
4–8 years Adults Older age 4–6 years: M/F Older age
Micronutrients 4–8 years Adults Older age Adults
9–13 years 19–50 years: 51 to >70 years: b 7–9 years: M/F 51 to >70 years:
9–13 years 19–50 years: 51 to >70 years 19–50 years: M/F
14–18 years: M/F b M/F 10–14 years: M/F M/F
14–18 years
Nutrients 2018, 10, 1531
195
30 150
Selenium, 27–41/26–58
40 55 55 280 400 400 36–73/31–54 39–62/34–55
μg/day 29–110/28–104
55 400
39–59/30–38
a Although adequate intake values are provided by the Institute of Medicine for infants (0–12 months) and recommended dietary allowances for children (1–3 years) [78], there are few
data regarding micronutrient deficiencies in this age groups in industrialized countries and these ages have therefore not been included in this table; b values differ in pregnancy and
lactation. F, females; M, males; ND, not determined.
Nutrients 2018, 10, 1531
196
Nutrients 2018, 10, 1531
vitamin C are unable to counteract the oxidative stress observed in pneumonia [104]. Increased
production of ROS during the immune response to pathogens may decrease vitamin C levels
further [105]. Vitamin D deficiency increases the risk of infection and autoimmune diseases such as
multiple sclerosis and diabetes, probably related to activity of vitamin D receptors, which are found
throughout the immune system [106,107].
Considering the importance of micronutrients in immunity, and the fact that many people of
all ages have single or multiple micronutrient deficiencies that can have detrimental immunological
effects, there is a rationale for micronutrient supplementation to restore concentrations to recommended
levels, especially after an infection, and to support immune function and maintenance. To avoid any
unwanted side effects, it is of course important to ensure that supplementation does not exceed
recommended tolerable upper intake levels (Table 2), the highest level of daily nutrient intake that
is likely to pose no risk of adverse health effects in most people [78]. Although this is theoretically
possible, the reported micronutrient intake data in Table 2 suggest that over-supplementation is
unlikely with most micronutrients, perhaps with the exception of vitamin A in children. It should be
noted that the safety margins in micronutrient supplements ensure that proper consumption does not
result in over-supplementation, and that food supplement labels should be carefully read to avoid
misuse and the potential for over-supplementation.
As no single biomarker exists that accurately reflects the effects of supplementation on the immune
response, clinical outcomes are instead used to determine the effectiveness of supplementation [49,69].
197
Table 3. Impact of micronutrient deficiency and supplementation on immune responses and the risk of infection.
198
Increased susceptibility to infections (e.g., diarrhea, RTI, measles,
Not beneficial in pneumonia [14]
malaria) [14,71]
Deficiency rare in humans [49]
Vitamin E Impairs both humoral and cell-mediated aspects of adaptive Older people: reduced RTI [71]
immunity, including B and T cell function [14]
Lymphocytopenia, reduced lymphoid tissue weight, reduced
Vitamin B6 responses to mitogens, general deficiencies in cell-mediated
immunity, lowered antibody responses [49]
Depressed immune responses (e.g., delayed-type hypersensitivity
Vitamin B12
response, T-cell proliferation) [49] *
Depressed immune responses (e.g., delayed-type hypersensitivity
Folate
response, T-cell proliferation) [49] *
Restoration of thymulin activity, increased numbers of cytotoxic T cells,
Decreased lymphocyte number and function, particularly T cells, reduced numbers of activated T helper cells (which can contribute to
increased thymic atrophy, altered cytokine production that autoimmunity), increased natural killer cell cytotoxicity, reduced
contributes to oxidative stress and inflammation [14] incidence of infections [14]
Zinc
Increased bacterial, viral and fungal infections (particularly diarrhea Children: reduction in duration of diarrhea and incidence of pneumonia
and pneumonia) [71] and diarrheal and respiratory morbidity [49] in at-risk children >6 month, but not in children 2–6 month [71];
Increased thymic atrophy and consequent risk of infection [97] reduced duration and severity of common cold symptoms [108];
improved outcomes in pneumonia, malaria and diarrheal symptoms [9]
Table 3. Cont.
199
Nutrients 2018, 10, 1531
200
Nutrients 2018, 10, 1531
lymphocytes, and induced a shift from memory T cells to naïve T cells [119]. Multiple micronutrient
supplementation in older people may also reduce antibiotic usage and lead to higher post-vaccination
immune responses [33].
Marginal zinc deficiency is common in older people, as their dietary intakes are generally lower and
plasma zinc concentrations decline with age, possibly connected to impaired absorption, alterations in
cellular uptake, and epigenetic dysregulation of DNA methylation or the methionine/transsulfuration
pathway, for example [14]. Supplementation with low to moderate doses of zinc in healthy older people
can help to restore thymulin activity, increase the numbers of cytotoxic T cells, reduce the number of
activated Th cells (which contribute to autoimmunity) and increase the cytotoxicity of NK cells [14],
immunological benefits that help to reduce the incidence of infections such as common cold, cold sores
and influenza [120], as well as the incidence and morbidity of pneumonia [121]. There are some reports
that an adequate zinc supply could prevent degenerative age-related diseases including infection
and cancer [122]. Sufficient vitamin C is also important in older people, who are at risk of vitamin
C deficiency, especially females [96]. Adequate vitamin C intakes can optimize cell and tissue levels
and help to protect against respiratory and systemic infections (e.g., reduced duration and severity of
pneumonia [71]), while higher levels are required during infection to compensate for the increased
inflammatory response and metabolic demand induced by the pathogen, and thus help to reduce
the duration and severity of symptoms [12]. Supplementation with vitamin E in older people has
been shown to significantly improve NK cytotoxic activity, neutrophil chemotaxis and the phagocytic
response, and enhance mitogen-induced lymphocyte proliferation and IL-2 production [123]. Vitamin E
can also improve T-cell-mediated immunity and increase the production of antibodies in response to
the hepatitis B and tetanus vaccines [124]. The risk of upper respiratory tract infections, especially
common cold, was significantly lower after vitamin E supplementation in nursing home residents,
although there was no apparent effect on lower respiratory tract infections [125]. However, not all
studies have reported beneficial effects on respiratory tract infections with vitamin E supplementation
in older people [14].
6. Conclusions
The immune system undergoes many changes over the life course—developing and maturing
during childhood, potentially achieving peak function in early adulthood, and gradually declining
in most people in older age (Figure 3). Distinct immune features are present during each life stage,
and specific factors differentially affect immune function, with a resulting difference in the type,
prevalence and severity of infections with age. A common factor throughout life is the need for an
adequate supply of micronutrients, which play key roles in supporting immune function. Multiple
micronutrient deficiencies are common throughout the world, with the likelihood increasing with
age. Tailored supplementation based on the specific needs of each age group may help to provide an
adequate basis for optimal immune function. The available clinical data suggest that micronutrient
supplementation can reduce the risk and severity of infection and support a faster recovery. However,
much more research is required into the effects of micronutrient supplementation on immune functions
and on clinical outcomes. Nevertheless, current knowledge regarding the importance of micronutrients
in immunity, the effects of micronutrient deficiencies on the risk and severity of infection, and the
worldwide prevalence of an inadequate micronutrient status form a sound basis for the use of a
targeted multiple micronutrient supplement to support immunity over a person’s lifetime.
201
Nutrients 2018, 10, 1531
202
Figure 3. Differences in immunity and nutrition over a lifetime. Ca, calcium; Cu, copper; Fe, iron; I, iodine; Ig, immunoglobulin; Mg, magnesium; NK, natural killer;
RTI, respiratory tract infections; Se, selenium; Th, T helper cell; Zn, zinc.
Nutrients 2018, 10, 1531
Author Contributions: S.M., A.P. and P.C.C. conceived and co-wrote the review, P.C.C. had primary responsibility
for the final content.
Funding: This research received no external funding.
Acknowledgments: A draft of this manuscript was prepared by a professional medical writer (Deborah Nock,
Medical WriteAway, UK; funded by Bayer Consumer Care AG), with subsequent full review and approval by
all authors.
Conflicts of Interest: S.M. and A.P. are employed by Bayer Consumer Care Ltd., a manufacturer of multivitamins.
P.C.C. has received funding, as a Key Opinion Leader, from Bayer Consumer Care Ltd.
References
1. Castelo-Branco, C.; Soveral, I. The immune system and aging: A review. Gynecol. Endocrinol. 2014, 30, 16–22.
[CrossRef] [PubMed]
2. Pandya, P.H.; Murray, M.E.; Pollok, K.E.; Renbarger, J.L. The immune system in cancer pathogenesis:
Potential therapeutic approaches. J. Immunol. Res. 2016, 2016, 4273943. [CrossRef] [PubMed]
3. Maggini, S.; Maldonado, P.; Cardim, P.; Fernandez Newball, C.; Sota Latino, E. Vitamins C., D and zinc:
Synergistic roles in immune function and infections. Vitam. Miner. 2017, 6, 167. [CrossRef]
4. Alpert, P. The role of vitamins and minerals on the immune system. Home Health Care Manag. Pract. 2017, 29,
199–202. [CrossRef]
5. Calder, P. Conference on ‘Transforming the nutrition landscape in Africa’. Plenary Session 1: Feeding the
immune system. Proc. Nutr. Soc. 2013, 72, 299–309. [CrossRef] [PubMed]
6. Watson, R.R.; Zibadi, S.; Preedy, V.R. Dietary Components and Immune Function; Springer Science & Business
Media: Berlin, Germany, 2010.
7. Biebinger, R.; Hurrell, R.F. 3—Vitamin and mineral fortification of foods. In Food Fortification and Supplementation;
Ottaway, P.B., Ed.; Woodhead Publishing: Cambridge, UK, 2008; pp. 27–40.
8. Schaefer, E. Micronutrient deficiency in women living in industrialized countries during the reproductive
years: Is there a basis for supplementation with multiple micronutrients? J. Nutr. Disord. Ther. 2016, 6, 199.
[CrossRef]
9. Wintergerst, E.; Maggini, S.; Hornig, D. Immune-enhancing role of vitamin C and zinc and effect on clinical
conditions. Ann. Nutr. Metab. 2006, 50, 85–94. [CrossRef] [PubMed]
10. Haryanto, B.; Suksmasari, T.; Wintergerst, E.; Maggini, S. Multivitamin supplementation supports immune
function and ameliorates conditions triggered by reduced air quality. Vitam. Miner. 2015, 4, 1–15.
11. Maggini, S.; Beveridge, S.; Sorbara, J.; Senatore, G. Feeding the immune system: The role of micronutrients
in restoring resistance to infections. CAB Rev. 2008, 3, 1–21. [CrossRef]
12. Carr, A.; Maggini, S. Vitamin C and immune function. Nutrients 2017, 9, 1211. [CrossRef] [PubMed]
13. Saeed, F.; Nadeem, M.; Ahmed, R.; Nadeem, M.; Arshad, M.; Ullah, A. Studying the impact of nutritional
immunology underlying the modulation of immune responses by nutritional compounds—A review.
Food Agric. Immunol. 2016, 27, 205–229. [CrossRef]
14. Micronutrient Information Center. Immunity in Depth. Available online: http://lpi.oregonstate.edu/mic/
health-disease/immunity (accessed on 17 April 2018).
15. Meydani, S.; Ribaya-Mercado, J.; Russell, R.; Sahyoun, N.; Morrow, F.; Gershoff, S. Vitamin B-6 deficiency
impairs interleukin 2 production and lymphocyte proliferation in elderly adults. Am. J. Clin. Nutr. 1991, 53,
1275–1280. [CrossRef] [PubMed]
16. Wintergerst, E.; Maggini, S.; Hornig, D. Contribution of selected vitamins and trace elements to immune
function. Nutr. Metab. 2007, 51, 301–323. [CrossRef] [PubMed]
17. World Health Organization; Food and Agricultural Organization of the United Nations. Part 2. Evaluating
the public health significance of micronutrient malnutrition. In Guidelines on Food Fortification with
Micronutrients; World Health Organization: Geneva, Switzerland, 2006.
18. Maggini, S.; Wintergerst, E.S.; Beveridge, S.; Hornig, D.H. Selected vitamins and trace elements support
immune function by strengthening epithelial barriers and cellular and humoral immune responses.
Br. J. Nutr. 2007, 98, S29–S35. [CrossRef] [PubMed]
19. Simon, A.; Hollander, G.; McMichael, A. Evolution of the immune system in humans from infancy to old
age. Proc. R. Soc. B 2015, 282, 20143085. [CrossRef] [PubMed]
203
Nutrients 2018, 10, 1531
20. Butler, E.; Kehrli, M. Immunoglobulins and immunocytes in the mammary gland and its secretions. In Mucosal
Immunology; Mestecky, J., Lamm, M., Ogra, P., Strober, W., Bienenstock, J., McGhee, J., Mayer, L., Eds.; Elsevier:
Amsterdam, The Netherland, 2005; pp. 1763–1793.
21. Field, C.J. The immunological components of human milk and their effect on immune development in
infants. J. Nutr. 2005, 135, 1–4. [CrossRef] [PubMed]
22. Witkowska-Zimny, M.; Kaminska-El-Hassan, E. Cells of human breast milk. Cell. Mol. Biol. Lett. 2017, 22, 11.
[CrossRef] [PubMed]
23. Brodin, P.; Davis, M. Human immune system variation. Nat. Rev. Immunol. 2017, 17, 21–29. [CrossRef]
[PubMed]
24. MacGillivray, D.; Kollmann, T. The role of environmental factors in modulating immune responses in early
life. Front. Immunol. 2014, 5, 434. [CrossRef] [PubMed]
25. Valiathan, R.; Ashman, M.; Asthana, D. Effects of ageing on the immune system: Infants to elderly. Scand. J.
Immunol. 2016, 83, 255–266. [CrossRef] [PubMed]
26. Hepworth, M.; Sonnenberg, G. Regulation of the adaptive immune system by innate lymphoid cells.
Curr. Opin. Immunol. 2014, 27, 75–82. [CrossRef] [PubMed]
27. Tortura, G.J.; Derrickson, B. Metabolism and nutrition. In Principles of Anatomy and Physiology, 8th ed.;
John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2014; pp. 940–978.
28. Wang, K.; Wei, G.; Liu, D. CD19: A biomarker for B cell development, lymphoma diagnosis and therapy.
Exp. Hematol. Oncol. 2012, 1, 36. [CrossRef] [PubMed]
29. Berg, T.; Johansson, S. Immunoglobulin levels during childhood, with special regard to IgE. Acta Paediatr.
1969, 58, 513–524. [CrossRef]
30. Pawelec, G. Hallmarks of human “immunosenescence”: Adaptation or dysregulation? Immun. Ageing 2012,
9, 15. [CrossRef] [PubMed]
31. Klein, S.; Hodgson, A.; Robinson, D. Mechanisms of sex disparities in influenza pathogenesis. J. Leukoc. Biol.
2012, 92, 67–73. [CrossRef] [PubMed]
32. Fulop, T.; Witkowski, J.; Pawelec, G.; Alan, C.; Larbi, A. On the immunological theory of aging. In Aging:
Facts and Theories; Robert, L., Fulop, T., Eds.; Karger: Basel, Switzerland, 2014; Volume 39, pp. 163–176.
33. Chandra, R. Nutrition and the immune system from birth to old age. Eur. J. Clin. Nutr. 2002, 56, S73–S76.
[CrossRef] [PubMed]
34. Montecino-Rodriguez, E.; Berent-Maoz, B.; Dorshkind, K. Causes, consequences, and reversal of immune
system aging. J. Clin. Investig. 2013, 123, 958–965. [CrossRef] [PubMed]
35. Ventura, M.; Casciaro, M.; Gangemi, S.; Buquicchio, R. Immunosenescence in aging: Between immune cells
depletion and cytokines up-regulation. Clin. Mol. Allergy 2017, 15, 21. [CrossRef] [PubMed]
36. Pawelec, G. Does the human immune system ever really become “senescent”? F1000Research 2017, 6, 1323.
[CrossRef] [PubMed]
37. Goronzy, J.; Weyand, C. Immune aging and autoimmunity. Cell. Mol. Life Sci. 2012, 69, 1615–1623. [CrossRef]
[PubMed]
38. Jafarzadeh, A.; Sadeghi, M.; Karam, G.A.; Vazirinejad, R. Salivary IgA and IgE levels in healthy subjects:
Relation to age and gender. Braz. Oral Res. 2010, 24, 21–27. [CrossRef] [PubMed]
39. Grewe, M. Chronological ageing and photoageing of dendritic cells. Clin. Exp. Dermatol. 2001, 26, 608–612.
[CrossRef] [PubMed]
40. Pand, A.A.; Qian, F.; Mohanty, S.; van Duin, D.; Newman, F.K.; Zhang, L.; Chen, S.; Towle, V.; Belshe, R.B.;
Fikrig, E.; et al. Age-associated decrease in TLR function in primary human dendritic cells predicts influenza
vaccine response. J. Immunol. 2010, 184, 2518–2527. [CrossRef] [PubMed]
41. Hemmi, H.; Akira, S. TLR signalling and the function of dendritic cells. Chem. Immunol. Allergy 2005, 86,
120–135. [PubMed]
42. Zhang, Y.; Wallace, D.; de Lara, C.; Ghattas, H.; Asquith, B.; Worth, A.; Griffin, G.; Taylor, G.; Tough, D.;
Beverley, P.; et al. In vivo kinetics of human natural killer cells: The effects of ageing and acute and chronic
viral infection. Immunology 2007, 121, 258–265. [CrossRef] [PubMed]
43. Hazeldine, J.; Lord, J. The impact of ageing on natural killer cell function and potential consequences for
health in older adults. Ageing Res. Rev. 2013, 12, 1069–1078. [CrossRef] [PubMed]
204
Nutrients 2018, 10, 1531
44. Fulop, T.; Larbi, A.; Dupuis, G.; Le Page, A.; Frost, E.; Cohen, A.; Witkowski, J.; Franceschi, C.
Immunosenescence and inflamm-aging as two sides of the same coin: Friends or foes? Front. Immunol. 2017,
8, 1960. [CrossRef] [PubMed]
45. Monto, A.; Ullman, B. Acute respiratory illness in an American community: The Tecumseh study. JAMA
1974, 227, 164–169. [CrossRef] [PubMed]
46. World Health Organization. Influenza (Seasonal). Fact Sheet. Available online: http://www.who.int/
mediacentre/factsheets/fs211/en/ (accessed on 28 April 2018).
47. National Institute for Health and Care Excellence Diarrhoea and Vomiting Caused by Gastroenteritis in
under 5s: Diagnosis and Management. Clinical Guideline [CG84]. Available online: https://www.nice.org.
uk/guidance/cg84 (accessed on 28 April 2018).
48. GBD 2015 LRI Collaborators. Estimates of the global, regional, and national morbidity, mortality,
and aetiologies of lower respiratory tract infections in 195 countries: A systematic analysis for the Global
Burden of Disease Study 2015. Lancet Infect. Dis. 2017, 17, 1133–1161. [CrossRef]
49. Calder, P.; Prescott, S.; Caplan, M. Scientific Review: The Role of Nutrients in Immune Function of Infants and
Young Children; Emerging Evidence for Long-Chain Polyunsaturated Fatty Acids; Mead Johnson & Company:
Glenview, IL, USA, 2007.
50. Bresnahan, K.; Tanumihardjo, S. Undernutrition, the acute phase response to infection, and its effects on
micronutrient status indicators. Adv. Nutr. 2014, 5, 702–711. [CrossRef] [PubMed]
51. Milner, J.; Beck, M. Micronutrients, immunology and inflammation. The impact of obesity on the immune
response to infection. Proc. Nutr. Soc. 2012, 71, 298–306. [CrossRef] [PubMed]
52. Gleeson, M. Effects of exercise on immune function. Sports Sci. Exch. 2015, 28, 1–6.
53. Gleeson, M. Immunological aspects of sports nutrition. Immunol. Cell Biol. 2016, 94, 117–123. [CrossRef]
[PubMed]
54. Nieman, D. Immunonutrition support for athletes. Nutr. Rev. 2008, 66, 310–320. [CrossRef] [PubMed]
55. Campbell, J.; Turner, J. Debunking the myth of exercise-induced immune suppression: Redefining the impact
of exercise on immunological health across the lifespan. Front. Immunol. 2018, 9, 648. [CrossRef] [PubMed]
56. Segerstrom, S.; Miller, G. Psychological stress and the human immune system: A meta-analytic study of
30 years of inquiry. Psychol. Bull. 2004, 130, 601–630. [CrossRef] [PubMed]
57. Romeo, J.; Wärnberg, J.; Nova, E.; Díaz, L.E.; Gómez-Martinez, S.; Marcos, A. Moderate alcohol consumption
and the immune system: A review. Br. J. Nutr. 2007, 98, S111–S115. [CrossRef] [PubMed]
58. Besedovsky, L.; Lange, T.; Born, J. Sleep and immune function. Eur. J. Physiol. 2012, 163, 121–137. [CrossRef]
[PubMed]
59. Cohen, S.; Tyrrell, D.; Smith, A. Psychological stress and susceptibility to the common cold. N. Engl. J. Med.
1991, 325, 606–612. [CrossRef] [PubMed]
60. Nieman, D. Exercise, upper respiratory tract infection, and the immune system. Med. Sci. Sports Exerc. 1994,
26, 128–139. [CrossRef] [PubMed]
61. Risk Management Solutions. Learning from the 2009 H1N1 Influenza Pandemic. RMS Special Report.
Available online: http://static.rms.com/email/documents/liferisks/reports/learning-from-the-2009-h1n1-
influenza-pandemic.pdf (accessed on 28 April 2018).
62. Marshall, J.A.; Bruggink, L.D. The dynamics of norovirus outbreak epidemics: Recent insights. Int. J. Environ.
Res. Public Health 2011, 8, 1141–1149. [CrossRef] [PubMed]
63. Man, S. The clinical importance of emerging Campylobacter species. Nat. Rev. Gastroenterol. Hepatol. 2011, 8,
669–685. [CrossRef] [PubMed]
64. Centers for Disease Control and Prevention. Norovirus Worldwide. Available online: https://www.cdc.
gov/norovirus/worldwide.html (accessed on 29 July 2018).
65. Yoshikawa, T. Epidemiology and unique aspects of aging and infectious diseases. Clin. Infect. Dis. 2000, 30,
931–933. [CrossRef] [PubMed]
66. Heikkinen, T.; Jarvinen, A. The common cold. Lancet 2003, 361, 51–59. [CrossRef]
67. Eccles, R. Mechanisms of symptoms of common cold and flu. In Common Cold; Eccles, R., Weber, O., Eds.;
Birkhauser Verlag: Basel, Switzerland, 2009; pp. 23–45.
68. Ballinger, M.; Standiford, T. Postinfluenza bacterial pneumonia: Host defenses gone awry. J. Interferon
Cytokine Res. 2010, 30, 643–652. [CrossRef] [PubMed]
205
Nutrients 2018, 10, 1531
69. Albers, R.; Bourdet-Sicard, R.; Braun, D.; Calder, P.; Herz, U.; Lambert, C.; Lenoir-Wijnkoop, I.; Meheurst, A.;
Ouwehand, A.; Phothirath, P.; et al. Monitoring immune modulation by nutrition in the general population:
Identifying and substantiating effects on human health. Br. J. Nutr. 2013, S110–S130. [CrossRef] [PubMed]
70. Bhaskaram, P. Micronutrient malnutrition, infection, and immunity: An overview. Nutr. Rev. 2002, 60,
S40–S45. [CrossRef] [PubMed]
71. Prentice, S. They are what you eat: Can nutritional factors during gestation and early infancy modulate the
neonatal immune response? Front. Immunol. 2017, 8, 1641. [CrossRef] [PubMed]
72. Butte, N.; Lopez-Alarcon, M.; Garza, C. Nutrient Adequacy of Exclusive Breastfeeding for the Term Infant
during the First Six Months of Life. Available online: http://apps.who.int/iris/handle/10665/42519
(accessed on 12 December 2017).
73. Björklund, K.; Vahter, M.; Palm, B.; Grandér, M.; Lignell, S.; Berglund, M. Metals and trace element
concentrations in breast milk of first time healthy mothers: A biological monitoring study. Environ. Health
2012, 11, 1–8. [CrossRef] [PubMed]
74. Hall Moran, V.; Lowe, N.; Crossland, N.; Berti, C.; Cetin, I.; Hermoso, M.; Koletzko, B.; Dykes, F. Nutritional
requirements during lactation. Towards European alignment of reference values: The EURRECA network.
Matern. Child Health 2010, 6, 39–54. [CrossRef] [PubMed]
75. Kominiarek, M.; Rajan, P. Nutrition recommendations in pregnancy and lactation. Med. Clin. N. Am. 2016,
100, 1199–1215. [CrossRef] [PubMed]
76. Dawodu, A.; Tsang, R. Maternal vitamin D status: Effect on milk vitamin D content and vitamin D status of
breastfeeding infants. Adv. Nutr. 2012, 3, 353–361. [CrossRef] [PubMed]
77. Semba, R. Impact of micronutrient deficiencies on immune function. In Micronutrient Deficiencies during the
Weaning Period and the First Years of Life, Proceedings of the 54th Nestlé Nutrition Workshop, Pediatric Program,
São Paulo, Brazil, 26–30 October 2003; Pettifor, J., Zlotkin, S., Eds.; Nestlé Nutrition Institute Workshop Series:
Lausanne, Switzerland, 2004.
78. Institute of Medicine. Dietary Reference Intakes for Calcium and Vitamin D; The National Academies Press:
Washington, DC, USA, 2011.
79. Marasinghe, E.; Chackrewarthy, S.; Abeysena, C.; Rajindrajith, S. Micronutrient status and its relationship
with nutritional status in preschool children in urban Sri Lanka. Asia Pac. J. Clin. Nutr. 2015, 24, 144–151.
[PubMed]
80. Luo, R.; Shi, Y.; Zhou, H.; Yue, A.; Zhang, L.; Sylvia, S.; Medina, A.; Rozelle, S. Micronutrient deficiencies
and developmental delays among infants: Evidence from a cross-sectional survey in rural China. BMJ Open
2015, 5, e008400. [CrossRef] [PubMed]
81. Özden, T.A.; Gökçay, G.; Cantez, M.S.; Durmaz, Ö.; İşsever, H.; Ömer, B.; Saner, G. Copper, zinc and iron
levels in infants and their mothers during the first year of life: A prospective study. BMC Pediatr. 2015, 15,
157. [CrossRef] [PubMed]
82. Jardim-Botelho, A.; Queiroz Gurgel, R.; Simeone Henriques, G.; Dos Santos, C.B.; Afonso Jordão, A.;
Nascimento Faro, F.; Silveira Souto, F.M.; Rodrigues Santos, A.P.; Eduardo Cuevas, L. Micronutrient
deficiencies in normal and overweight infants in a low socio-economic population in north-east Brazil.
Paediatr. Int. Child Health 2016, 36, 198–202. [CrossRef] [PubMed]
83. Bailey, R.; West, K.J.; Black, R. The epidemiology of global micronutrient deficiencies. Ann. Nutr. Metab.
2015, 66, 22–33. [CrossRef] [PubMed]
84. World Health Organization. Food and Agricultural Organization of the United Nations. Part I. The role
of food fortification in the control of micronutrient malnutrition. In Guidelines on Food Fortification with
Micronutrients; Allen, L., de Benoist, B., Dary, O., Hurrell, R., Eds.; World Health Organization: Geneva,
Switzerland, 2006.
85. Mackey, A.; Picciano, M. Maternal folate status during extended lactation and the effect of supplemental
folic acid. Am. J. Clin. Nutr. 1999, 69, 285–292. [CrossRef] [PubMed]
86. Houghton, L.; Sherwood, K.; O’Connor, D. How well do blood folate concentrations predict dietary folate
intakes in a sample of Canadian lactating women exposed to high levels of folate? An observational study.
BMC Pregnancy Childbirth 2007, 7, 1–8. [CrossRef] [PubMed]
87. Gellert, S.; Ströhle, A.; Hahn, A. Breastfeeding woman are at higher risk of vitamin D deficiency than
nonbreastfeeding women - insights from the German VitaMinFemin study. Int. Breastfeed. J. 2017, 12, 1–10.
[CrossRef] [PubMed]
206
Nutrients 2018, 10, 1531
88. Milman, N.; Hvas, A.-M.; Bergholt, T. Vitamin D status during normal pregnancy and postpartum.
A longitudinal study in 141 Danish women. J. Perinat. Med. 2012, 40, 57–61. [CrossRef] [PubMed]
89. Dawodu, A.; Zalla, L.; Woo, J.G.; Herbers, P.M.; Davidson, B.S.; Heubi, J.E.; Morrow, A.L. Heightened
attention to supplementation is needed to improve the vitamin D status of breastfeeding mothers and infants
when sunshine exposure is restricted. Matern. Child Health 2014, 10, 383–397. [CrossRef] [PubMed]
90. Hannan, M.; Faraji, B.; Tanguma, J.; Longoria, N.; Rodriguez, R. Maternal milk concentration of zinc, iron,
selenium, and iodine and its relationship to dietary intakes. Biol. Trace Elem. Res. 2009, 127, 6–15. [CrossRef]
[PubMed]
91. Charlton, K.; Yeatman, H.; Lucas, C.; Axford, S.; Gemming, L.; Houweling, F.; Goodfellow, A.; Ma, G. Poor
knowledge and practices related to iodine nutrition during pregnancy and lactation in australian women:
Pre- and post-iodine fortification. Nutrients 2012, 4, 1317–1327. [CrossRef] [PubMed]
92. Mulrine, H.M.; Skeaff, S.A.; Ferguson, E.L.; Gray, A.R.; Valeix, P. Breast-milk iodine concentration declines
over the first 6 mo postpartum in iodine-deficient women. Am. J. Clin. Nutr. 2010, 92, 849–856. [CrossRef]
[PubMed]
93. Valent, F.; Horvat, M.; Mazej, D.; Stibilj, V.; Barbone, F. Maternal diet and selenium concentration in human
milk from an Italian population. J. Epidemiol. 2011, 21, 285–292. [CrossRef] [PubMed]
94. Brenna, J.T.; Varamini, B.; Jensen, R.G.; Diersen-Schade, D.A.; Boettcher, J.A.; Arterburn, L.M. Docosahexaenoic
and arachidonic acid concentrations in human breast milk worldwide. Am. J. Clin. Nutr. 2007, 85, 1457–1464.
[CrossRef] [PubMed]
95. Podzolkova, N.; Schaefer, E. Micronutrient intakes and status are frequently insufficient in breastfeeding
women. BMC Pregnancy Childbirth 2018, submitted for publication.
96. Elmadfa, I.; Meyer, A.; Nowak, V.; Hasenegger, V.; Putz, P.; Verstraeten, R.; Remaut-DeWinter, A.M.;
Kolsteren, P.; Dostálová, J.; Dlouhý, P.; et al. European Nutrition and Health Report. Forum Nutr. 2009, 62,
1–405. [PubMed]
97. Savino, W.; Dardenne, M. Nutritional imbalances and infections affect the thymus: Consequences on
T-cell-mediated immune responses. Proc. Nutr. Soc. 2010, 69, 636–643. [CrossRef] [PubMed]
98. Palacios, C.; Gonzalez, L. Is vitamin D deficiency a major global public health problem? J. Steroid Biochem.
Mol. Biol. 2014, 144PA, 138–145. [CrossRef] [PubMed]
99. Montgomery, S.; Streit, S.; Beebe, M.; Maxwell IV, P. Micronutrient needs of the elderly. Nutr. Clin. Pract.
2014, 29, 435–444. [CrossRef] [PubMed]
100. Drenowski, A.; Shultz, J. Impact of aging on eating behaviors, food choices, nutrition, and health status.
J. Nutr. Health Aging 2001, 5, 75–79.
101. High, K. Nutritional strategies to boost immunity and prevent infection in elderly individuals. Clin. Infect. Dis.
2001, 33, 1892–1900. [CrossRef] [PubMed]
102. Wiacek, M.; Zubrzycki, I.; Bojke, O.; Kim, H. Menopause and age-driven changes in blood level of fat- and
water-soluble vitamins. Climacteric 2013, 16, 689–699. [CrossRef] [PubMed]
103. Karaouzenea, N.; Merzouka, H.; Aribib, M.; Merzoukc, S.; Yahia Berrouiguet, A.; Tessiere, C.; Narce, M. Effects
of the association of aging and obesity on lipids, lipoproteins and oxidative stress biomarkers: A comparison of
older with young men. Nutr. Metab. Cardiovasc. Dis. 2011, 21, 792–799. [CrossRef] [PubMed]
104. Hemilä, H. Vitamin C and infections. Nutrients 2017, 9, 339. [CrossRef] [PubMed]
105. Hemilä, H.; Chalker, E. Vitamin C for preventing and treating the common cold. Cochrane Database Syst. Rev.
2013. Available online: https://www.cochranelibrary.com/cdsr/doi/10.1002/14651858.CD000980.pub4/
full (accessed on 17 October 2018).
106. Aranow, C. Vitamin D and the immune system. J. Investig. Med. 2011, 59, 881–886. [CrossRef] [PubMed]
107. Mangin, M.; Sinha, R.; Fincher, K. Inflammation and vitamin D: The infection connection. Inflamm. Res. 2014,
63, 803–819. [CrossRef] [PubMed]
108. Hemilä, H. Zinc lozenges may shorten the duration of colds: A systematic review. Open Respir. Med. J. 2011,
5, 51–58. [PubMed]
109. Johnston, C.; Barkyoumb, G.M.; Schumacher, S.S. Vitamin C supplementation slightly improves physical
activity levels and reduces cold incidence in men with marginal vitamin C status: A randomized controlled
trial. Nutrients 2014, 6, 2572–2583. [CrossRef] [PubMed]
110. Maggini, S.; Beveridge, S.; Suter, M. A combination of high-dose vitamin C plus zinc for the common cold.
J. Int. Med. Res. 2012, 40, 28–42. [CrossRef] [PubMed]
207
Nutrients 2018, 10, 1531
111. Graf, P.; Eccles, R.; Chen, S. Efficacy and safety of intranasal xylometazoline and ipratropium in patients
with common cold. Expert Opin. Pharmacother. 2009, 10, 889–908. [CrossRef] [PubMed]
112. Yamshchikov, A.; Desai, N.; Blumberg, H.; Ziegler, T.; Tangpricha, V. Vitamin D for treatment and prevention
of infectious diseases: A systematic review of randomized controlled trials. Endocr. Pract. 2009, 15, 438–449.
[CrossRef] [PubMed]
113. Charan, J.; Goyal, J.; Saxena, D.; Yadav, P. Vitamin D for prevention of respiratory tract infections:
A systematic review and meta-analysis. J. Pharmacol. Pharmacother. 2012, 3, 300–303. [CrossRef] [PubMed]
114. Bergman, P.; Lindh, Å.; Björkhem-Bergman, L.; Lindh, J. Vitamin D and respiratory tract infections:
A systematic review and meta-analysis of randomized controlled trials. PLoS ONE 2013, 8, e65835. [CrossRef]
[PubMed]
115. Martineau, A.; Jolliffe, D.; Hooper, R.; Greenberg, L.; Aloia, J.; Bergman, P.; Dubnov-Raz, G.; Esposito, S.;
Ganmaa, D.; Ginde, A.A.; et al. Vitamin D supplementation to prevent acute respiratory infections:
Systematic review and meta-analysis of individual participant data. BMJ 2017, 356, i6583. [CrossRef]
[PubMed]
116. Hamer, D.; Sempértegui, F.; Estrella, B.; Tucker, K.; Rodríguez, A.; Egas, J.; Dallal, G.; Selhub, J.; Griffiths, J.;
Meydani, S. Micronutrient deficiencies are associated with impaired immune response and higher burden of
respiratory infections in elderly Ecuadorians. J. Nutr. 2009, 139, 113–119. [CrossRef] [PubMed]
117. Penn, N.D.; Purkins, L.; Kelleher, J.; Heatley, R.V.; Mascie-Taylor, B.H.; Belfield, P.W. The effect of dietary
supplementation with vitamins A, C, and E on cell-mediated immune function in elderly log-stay patients:
A randomized controlled trial. Age Ageing 1991, 20, 169–174. [CrossRef] [PubMed]
118. Chandra, R. Effect of vitamin and trace-element supplementation on immune responses and infection in
elderly subjects. Lancet 1992, 340, 1124–1127. [CrossRef]
119. Schmoranzer, F.; Fuchs, N.; Markolin, G.; Carlin, E.; Sakr, L.; Sommeregger, U. Influence of a complex
micronutrient supplement on the immune status of elderly individuals. Int. J. Vitam. Nutr. Res. 2009, 79,
308–318. [CrossRef] [PubMed]
120. Prasad, A. Zinc: Mechanisms of host defense. J. Nutr. 2007, 137, 1345–1349. [CrossRef] [PubMed]
121. Meydani, S.; Barnett, J.; Dallal, G.; Fine, B.; Jacques, P.; et al. Serum zinc and pneumonia in nursing home
elderly. Am. J. Clin. Nutr. 2007, 86, 1167–1173. [CrossRef] [PubMed]
122. Mocchegiani, E.; Romeo, J.; Malavolta, M.; Costarelli, L.; Giacconi, R.; Diaz, L.; Marcos, A. Zinc: Dietary
intake and impact of supplementation on immune function in elderly. Age (Dordr.) 2013, 35, 839–860.
[CrossRef] [PubMed]
123. De la Fuente, M.; Hernanz, A.; Guayerbas, N.; Victor, V.; Arnalich, F. Vitamin E ingestion improves several
immune functions in elderly men and women. Free Radic. Res. 2008, 42, 272–280. [CrossRef] [PubMed]
124. Meydani, S.; Meydani, M.; Blumberg, J.; Leka, L.; Siber, G.; Loszewski, R.; Thompson, C.; Pedrosa, M.;
Diamond, R.; Stollar, B. Vitamin E supplementation and in vivo immune response in healthy elderly subjects.
A randomized controlled trial. JAMA 1997, 277, 1380–1386. [CrossRef] [PubMed]
125. Meydani, S.; Leka, L.; Fine, B.; Dallal, G.; Keusch, G.; Singh, M.; Hamer, D. Vitamin E and respiratory tract
infections in elderly nursing home residents: A randomized controlled trial. JAMA 2004, 292, 828–836.
[CrossRef] [PubMed]
© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
208
nutrients
Review
Selenium, Selenoproteins, and Immunity
Joseph C. Avery and Peter R. Hoffmann *
Department of Cell and Molecular Biology, John A. Burns School of Medicine, University of Hawaii,
651 Ilalo Street, Honolulu, HI 96813, USA; jcavery@hawaii.edu
* Correspondence: peterrh@hawaii.edu; Tel.: +1-808-692-1510; Fax: +808-692-1968
Abstract: Selenium is an essential micronutrient that plays a crucial role in development and a wide
variety of physiological processes including effect immune responses. The immune system relies
on adequate dietary selenium intake and this nutrient exerts its biological effects mostly through
its incorporation into selenoproteins. The selenoproteome contains 25 members in humans that
exhibit a wide variety of functions. The development of high-throughput omic approaches and novel
bioinformatics tools has led to new insights regarding the effects of selenium and selenoproteins
in human immuno-biology. Equally important are the innovative experimental systems that have
emerged to interrogate molecular mechanisms underlying those effects. This review presents a
summary of the current understanding of the role of selenium and selenoproteins in regulating
immune cell functions and how dysregulation of these processes may lead to inflammation or
immune-related diseases.
1. Introduction
Selenium was discovered by the Swedish chemist Jöns Jakob Berzelius in 1817 and was considered
a toxic element for humans and livestock for nearly 150 years [1]. However, in 1957, the benefits
of selenium for humans and other mammals were revealed in landmark studies by Klaus Schwartz
and Calvin Foltz who demonstrated that dietary selenium protected rats against liver necrosis [2].
Since then, the role of selenium as a trace mineral nutrient in human health and the mechanisms by
which it exerts its biological effects have become better understood. Adequate levels of bioavailable
selenium are functionally important for several aspects of human biology including the central nervous
system, the male reproductive biology, the endocrine system, muscle function, the cardiovascular
system, and immunity [3,4]. Many pathological conditions involving the immune system can be
affected by the selenium status in an individual, which can be influenced by several factors such as
the levels and forms of selenium ingested, the conversion of selenium compounds into metabolites,
and genetic characteristics that can impact the use of these metabolites. Selenium deficiency is rare in
the United States and Canada [5], but regions of China, New Zealand, and parts of Europe and Russia
have low levels of selenium in soil and food [6]. The extent to which immune-related diseases are
impacted by differences in selenium intake and how supplementation approaches may be utilized to
mitigate these health issues is not entirely clear. However, the development of new high-throughput
omic approaches and bioinformatics tools have improved our understanding of the effects of selenium
immuno-biology in humans. Additionally, novel experimental systems have provided valuable insight
into mechanisms underlying those effects.
The U.S. recommended dietary allowance for selenium for adults is 55 μg/day and most
individuals achieve this level while several other countries have higher recommended allowances
due to a lower average selenium status in their populations [7]. For example, adults in the U.K.
are recommended to ingest 60 μg/day for adult women and 75 μg/day for lactating women and
adult men [8]. Commonly used measures of a selenium status include plasma and serum selenium
concentrations as well as selenoprotein P levels and glutathione peroxidase activity [9,10]. The average
plasma selenium concentration in the U.S. is 70 ng/mL, which is relatively high with selenium
intake found to be lower in areas within China and Europe, in New Zealand, and in other parts of
the world [11,12]. Dietary selenium is obtained through a wide variety of foods including grains,
vegetables, seafood, meat, dairy products, and nuts [13]. The predominant form of selenium ingested
by humans is selenomethionine. However, other forms of selenium are also present in foods. Selenium
gets metabolized into various small molecular weight seleno-compounds including some that may
exert biological effects through redox reactions that can affect cellular processes like DNA repair and
epigenetics [14,15]. These bioactive metabolites include hydrogen selenide and methylated selenium
compounds like methylseleninic acid, which exerts chemo-preventive effects [16]. Most of the effects
of dietary selenium on immune functions are attributable to the insertion of this element into a
family of proteins called seleno-proteins. What separates selenium from other nutritional elements
is the fact that it is incorporated directly into proteins as the 21st amino acid, selenocysteine (Sec).
The synthesis of selenoproteins within cells requires a dedicated set of protein and tRNA factors
assembled on ribosomes along with the selenoprotein mRNA, which contains unique structural
elements. The coordinated interaction of these elements leads to co-translational insertion of Sec into
the nascent polypeptide when the ribosome encounters a uridine-guanosine-adenosine (UGA) codon,
which is typically used as a stop codon in other mRNAs [17]. Under conditions of low selenium status,
this translational process stalls at the UGA codon and both the mRNA and truncated protein may get
degraded through two separate processes called nonsense-mediated decay (NMD) and destruction via
C-end degrons (DesCEND), respectively [18,19]. Certain mRNA characteristics potentially play a role
in NMD sensitivity such as the location of the Sec codon (UGA) relative to exon–exon junctions [18].
Therefore, the selenium status is directly related to levels of different selenoproteins in different
tissues. Given the combined effects of NMD and DesCEND, there appears to be a hierarchy of
selenoprotein synthesis that results in some family members having a higher priority of expression
under selenium-limiting conditions [20]. In addition, certain tissues like the brain, endocrine tissues,
and testes retain selenium under deficient conditions shed light on the priorities given to different
physiological systems when selenium levels are low.
In humans, 25 selenoproteins have been identified and 24 of those exist as Sec-containing
proteins in rodents [21], which highlights the value of rodent models for determining roles for
members of this protein family in immune responses. Selenoproteins exhibit a wide variety of
tissue distribution and functions [17]. While many members of the selenoprotein family function
as enzymes involved in redox reactions, some are likely not enzymes themselves and functions are
gradually becoming better understood for these non-enzymatic members. The most completely
characterized selenoprotein enzymes related to immune functions include glutathione peroxidases
(GPXs), thioredoxin reductases (TXNRDs), iodothyronine deiodinases (DIOs), methionine-R-sulfoxide
reductase B1 (MSRB1), and selenophosphate synthetase 2 (SPS2). For non-enzymatic selenoproteins,
the best characterized in terms of immune cell function is selenoprotein K (SELENOK). Table 1 lists
selenoproteins and their functions and a more detailed discussion of roles for individual selenoproteins
in different immune cells and tissues is provided below.
210
Table 1. Summary of Selenoprotein Functions.
211
uridine-guanosine-guanosine-thymodine (UGGT) and improves protein quality control by
Selenoprotein F SELENOF, Selenoprotein 15, SEP15 correcting misglycosylated/misfolded glycoproteins via the calnexin-calreticulin- endoplasmic
reticulum proten 57 (ERp57) axis and pH-dependent endoplasmic reticulum proten 44 (ERp44
)system [37,38].
Selenoprotein H SELENOH, SELH, C11orf31 Nuclear localization, which is involved in redox sensing and transcription [39,40].
Selenoprotein I SELENOI, SELI, EPT1 Involved in phospholipid biosynthesis [41].
Transmembrane protein localized to the endoplasmic reticulum (ER) and involved in calcium
Selenoprotein K SELENOK, SELK
flux in immune cells and ER associated degradation in cell lines [42,43].
Thioredoxin-like ER-resident protein that may be involved in the regulation of body weight and
Selenoprotein M SELENOM, SELM, SEPM
energy metabolism [44].
Transmembrane protein localized to ER. Mutations lead to multiminicore disease and other
Selenoprotein N SELENON, SELN, SEPN1
myopathies [45,46].
Mitochondrial protein that contains a C-X-X-U (where C is cytosine, X is any nucleotide, and U
Selenoprotein O SELENOO, SELO
is uridine) motif suggestive of the redox function [47].
Selenoprotein P SELENOP, SEPP1, SeP, SELP, SEPP Secreted into plasma for selenium transport to tissues [20,48].
Selenoprotein S SELENOS, SELS, SEPS1, VIMP Transmembrane protein found in ER involved in ER associated degradation [49,50].
Oxidoreductase localized to the Golgi complex and ER and manifests a thioredoxin-like fold
Selenoprotein T SELENOT, SELT and is involved in redox regulation and cell anchorage. Complexes with UGGTs to improve
PQC. Deficiency leads to early embryonic lethality [51].
Selenoprotein V SELENOV, SELV Testes-specific expression [21].
Selenoprotein W SELENOW, SELW, SEPW1 Putative antioxidant role, which may be important in muscle growth [52].
Selenophosphate synthetase 2 SEPHS2, SPS2 Involved in synthesis of all selenoproteins including itself [53].
Nutrients 2018, 10, 1203
As mentioned above, nearly all tissues are affected by changes in the selenium status or
selenoprotein expression. While the focus of this review is on the immune system, it is important
to first touch on other physiological systems impacted by the levels of selenium and selenoproteins.
Embryonic lethality arising from deletion of the trsp gene encoding the Sec-tRNA required for
translation [54] demonstrates the essential nature of selenoproteins. In fact, there have been four
individual selenoprotein knockout mice in which gene ablation was shown to result in embryonic
lethality: GPX4, TXNRD1 and 2, and Selenoprotein T (SELENOT) [32,51,55,56]. An essential role
for one of these selenoproteins in the area of development was demonstrated by the recent study,
which showed that GPX4 protects a critical population of interneurons from ferroptotic cell death [29].
In the muscular system, genetic maladies involving selenoproteins include multi-minicore diseases
(MmD) such as rigid spine syndrome (RSS) resulting from mutations in the human gene encoding
Selenoprotein N (SELENON) [57,58] and an associative dysfunction of the ryanodine receptor 1 (RyR1)
receptor [59]. Transgenic overexpression of some selenoproteins potentially regenerates wasted muscle
in mice [60]. Thyroid hormone metabolism is dependent upon the combined actions of the three
selenoproteins known as iodothyronine deiodinases 1-3 (DIO1-3) [61]. Thus, selenium deficiencies
can affect thyroid gland function and the many physiological systems impacted by thyroid hormone
activity. In the hepatic system, selenium is absorbed from the gastrointestinal tract and utilized for
biosynthesis of selenoproteins including Selenoprotein P (SELENOP), which is the primary plasma
selenium transport protein [62]. Several groups have observed that SELENOP inactivation results
in normal hepatic selenium levels while selenium content in other tissues decreases significantly.
This reduces the total GPX and TXNRD pools [63,64]. Consequently, those organs that rely on
SELENOP-mediated selenium delivery become deficient when some tissues are given ‘priority’ over
others for retention of this element since delivery through SELENOP decreases.
The central nervous system is appreciably dependent on an adequate selenium supply and,
as mentioned above, diets that are slightly deficient in selenium do not elicit neurological deficits
due to the preservation of selenium content in the central nervous tissue during dietary selenium
restriction [65]. On the other hand, a targeted reduction in brain selenium reduces SELENOP
bioavailability and causes spontaneous neurological deficits [66], which are reversed by selenium
supplementation [67,68]. Additionally, overexpression of TRX1 has been found to mitigate oxidative
challenges in the brain [69]. GPX1 was the first mammalian selenoprotein to be discovered [70,71]
and has been shown to protect the brain from oxidative insults. Like GPX1, GPX4 protects cortical
neurons from exogenous oxidative stress-inducing agents [72,73]. Importantly, the protein oxidation
product methionine-R-sulfoxide contributes to neurodegenerative diseases and can be repaired
by thioredoxin-dependent selenoenzyme MSRB1, which reduces methionine-R-sulfoxide back to
methionine [36]. Inactivation of MSRB1, however, does not produce neurological deficits [36]. In the
kidney, several studies have identified the expression of DIOs, thioredoxin reductases (TRs), and GPXs,
but their respective roles have not been fully elucidated. Burk et al. demonstrated that glutathione
(GSH) deprivation causes severe pathogenic nephropathy [74] while podocyte-specific ablation of the
trsp gene in diabetic mice did not enhance markers of nephropathic disease. Moreover, murine renal
expression of GPX1 has been reported not to be protective against diabetic nephropathy.
For clinically diagnosed disorders, Keshan disease (KD) is perhaps the most firmly established
selenium deficiency-based pathology. This cardiomyopathy was first described in rural areas
of China due to low selenium content in foods [75]. There is evidence in mouse models that
selenium deficiency promotes the conversion of nonvirulent coxsackievirus B3 strains into a more
virulent strain due to an increased oxidative stress [76], which suggests that this infectious agent
may be a cofactor. Selenoprotein deficiency may also promote osteochondral diseases including
Kashin-Beck disease (KBD). This disease is a poly-pathogenic, degenerative osteochondropathy leading
to chondrocyte necrosis [77] and apoptosis [78–80], which results in growth retardation and secondary
osteoarthrosis [81]. KBD is mainly endemic to Tibet, China, Siberia, and North Korea and is caused in
part by poor selenium levels in soil that usually affects children between the ages of 5 to 15 [81,82].
212
Nutrients 2018, 10, 1203
In 1998, Moreno-Reyes et al. established the relationship between this osteoarthropathy and selenium
deficiency in rural Tibet [82].
213
Nutrients 2018, 10, 1203
3. Leukocyte Functions
Adaptive immunity is affected by selenium intake including the activation and functions of T
and B cells. One immunological feature of selenium levels in vivo is the positive effect that higher
selenium has on the proliferation and differentiation of cluster of differentiation(CD)4+ T helper
(Th) cells. There are several reports of the skewing of T cell immunity toward Th1 phenotypes.
For example, our laboratory used a mouse model of viral antigen vaccination to test effects of low
(0.087 ppm), medium (0.25 ppm), and high (1.0 ppm) selenium diets and found that Th1 immunity was
enhanced along with the T cell receptor signal strength [101]. In a separate study, oral administration
of synthetic selenium nanoparticles induced a robust Th1 cytokine pattern after a hepatitis B surface
antigen vaccination in a mouse model [102]. Less information is available regarding the effects of
selenium on cytotoxic CD8+ T cells even though cytotoxic T cells from aged mice (24 months old)
showed enhanced mitogen-induced proliferation when treated with selenium supplementation [103].
Mouse knockout models have shown roles for selenoproteins in antibody production. In particular,
T cell deletion of the trsp gene responsible for the synthesis of all selenoproteins not only affected T cell
maturation and activation but reduced the T cell ‘help’ provided to B cells for secreting antibodies,
which is determined by low levels of serum immunoglobulin [98]. A small study in humans showed a
positive effect on antibody titers against the diphtheria vaccine with selenium supplementation that
correspond to increased lymphocyte counts [104]. In a more recent study involving Selenoprotein F
(SELENOF) knockout mice, elevated levels of immunoglobulins were detected in the sera that were
nonfunctional [38]. The authors of this study concluded that SELENOF functions as a gatekeeper of
immunoglobulins in the endoplasmic reticulum (ER), which supports the redox quality control of
these proteins and likely other proteins.
Innate immune cell functions have also been shown to be impacted by selenium levels.
Macrophages are affected by selenium levels in terms of their inflammatory signaling capacity
and anti-pathogen activities. Activation of macrophages through pathogen-associated molecular
patterns like lipopolysaccharide (LPS) generates an oxidative burst. Additionally, macrophage
activation involves the release of cytokine mediators and arachidonic acid-derived prostaglandins
like prostaglandin E2 (PGE2), thromboxane A2 (TXA2), and prostaglandin D2 (PGD2) as well as its
metabolite 15-Deoxy-Delta-12,14-prostaglandin J2 (15d-PGJ2). It was shown that selenium induces a
phenotypic switch in macrophage activation from a classically activated, pro-inflammatory phenotype
(M1) toward an alternatively activated, anti-inflammatory phenotype (M2) [105]. Regarding the latter
214
Nutrients 2018, 10, 1203
215
Nutrients 2018, 10, 1203
effects on the immune system were not determined [127]. However, several researchers have pointed
out that some trace nutrients that may be used as supplements to restore immunity and lung function
may also be exploited by M. tuberculosis to promote growth of the pathogen [128]. This has been
supported by data involving the growth of this bacteria under different selenium concentrations [129].
The beneficial effects of a higher selenium status have been supported for some viral infections
even though there are some studies that do not conclusively demonstrate effective improvements in
anti-viral immunity [130]. Moreover, the antioxidant properties of some selenoproteins have been
suggested to contribute to boosting anti-viral immunity [131]. However, some selenoproteins that
have not been established as antioxidant enzymes like SELENOK can also play key roles in protecting
against viruses [42]. The chronic hepatitis C virus (HCV) has been shown to influence oxidative stress
levels in humans and an association between HCV load and the selenium status has been associated
with a documented selenium status [132]. Oxidative stress can have genomic altering effects on RNA
viruses that can lead to higher virulence of certain viruses themselves and this has been shown to
involve the selenium status in the case of coxsackievirus B3 [133]. Thus, the effects of selenium on
the virus in some cases may compound the influence of this micronutrient on the immune system.
Targeting individuals with low selenium intake or the elderly with a declining selenium status with
selenium supplementation may be an effective public health initiative for increasing vaccine responses
to viruses [134]. There is evidence to support a positive effect on adaptive immune responses to
vaccination against viral pathogens. This causes polio and influenza in populations with a low baseline
selenium status [93,135].
The most compelling data available regarding the role of selenium in anti-viral immunity are
those related to HIV infection, which is a global pandemic that particularly afflicts persons with
inadequate nutrition and directly impairs immunity [136]. Selenium is one micronutrient implicated
in disease progression. Low selenium intake has been associated with HIV prevalence [137,138] and
the status of CD4+ T cell numbers has been correlated with selenium levels in HIV+ patients [139].
There is some evidence that selenium malabsorption or overutilization in Acquired immune deficiency
syndrome (AIDS) patients may affect or be affected by disease progression [140,141]. In particular,
selenium-deficient HIV+ patients tend to present with disrupted hemodynamics such as depressed
selenium plasma and erythrocyte levels, diminished glutathione peroxidase activity, and stunted
cardiac selenium bioavailability. A plasma selenium status is conventionally assessed by SELENOP
levels and GPX activity as well as selenium levels, which respond differently to changes in selenium
consumption [142]. Thus, it is difficult to directly compare studies using different selenium status
readouts [143]. Anti-retroviral therapies may also confound the selenium status [144,145]. However,
some studies have not supported this notion [146]. Additionally, selenium is often combined with other
nutrients for intervention studies, which makes assessment of its impact difficult to distinguish from
other nutritional components. Several cohort studies have illustrated an association between selenium
deficiency and progression to AIDS-related mortality [147]. Remarkably, randomized controlled trials
demonstrated that selenium supplementation minimized hospitalizations and diarrheal morbidity and
improved CD4+ T cell counts [141,148]. Similarly, an inhibitory effect of selenium on HIV in vitro due to
the radical scavenger effects of glutathione peroxidase has been reported [141]. Glutathione peroxidase
and other antioxidant selenoenzymes along with catalase have been implicated in decreasing a viral
activation impact on redox control [141,149]. Thus, the potential benefits of selenium supplementation
for HIV infection likely resides in the redox regulating selenoenzymes and resides less with the
pro-oxidant seleno-metabolites that are found to affect cancer.
216
Nutrients 2018, 10, 1203
status was not found to be a factor in cancer progression in a number of other studies [153–156].
From the perspective of research in humans, it has proven difficult to separate the direct effects that
selenium has on carcinogenesis from its impact on the growth of established tumors as well as its
influence on cancer immunity. One of the direct anti-cancer effects of selenium is related to the ability
of seleno-compounds to induce oxidative stress and DNA damage accumulation and, consequently,
apoptosis [15]. Other direct effects of selenium on established tumors in humans are less clear and
this is particularly true for those effects that are exerted through the immune system. For example,
there is some evidence from one human study suggesting an inhibitory effect of selenium on the
epithelial-to-mesenchymal transition (EMT) that drives metastasis [157]. This was accompanied by the
capacity of higher levels of selenium to down-regulate expression of genes involved in wound healing
and inflammation, which are both related to EMT. The idea that selenium supplementation may be
used to support the immune system during cancer treatment has been supported by some studies
including those related to childhood leukemia and neutropenia [158,159]. Intervention studies showed
positive effects of selenium on mitigating neutropenia in children suffering from leukemia/lymphomas
as well as solid tumors [160].
There is evidence that GPX4 modulates hepatocellular carcinoma (HCC) in both humans and
rodent models. In humans, GPX4 expression in tumors positively correlated with patient survival and
was linked to pathways that regulate cell proliferation, motility, tissue remodeling, and immune
responses with a particular effect on M1 macrophage polarization [161]. Corroborative results
demonstrate that overexpression of GPX4 decreased the growth of human HCC cell lines using
xenotransplantation into immune-deficient non-diabetic (NOD) mice. These findings are consistent
with previous studies showing that inhibition of GPX4 expression by siRNA in HCC cells increased
the formation of Vascular endothelial growth factor (VEGF) and IL-8 cytokines [162], which are both
clinically relevant adverse prognostic factors in HCC patients [162,163]. However, since NOD mice do
not include a competent immune system, it is difficult to interpret how the immune relevant data from
the human gene arrays can be related to the rodent studies.
The polarization of tumor-associated macrophages away from tolerogenic phenotypes and toward
anti-tumor M1 macrophages suggested in the above experiments with GPX4 overexpression was also
supported in selenium nanoparticle studies [164]. However, how selenium levels affect macrophage
polarization in the tumor microenvironment in human cancers remains to be determined. As discussed
in a previous review [165], higher levels of selenium can increase NK cell activity by preventing the
non-enzymatic formation of parafibrin that surrounds tumor cells and hinders immune surveillance
and by activating the NK cell population in the tumor microenvironment. The anti-tumoral activity of
NK cells requires the expression of the activating receptor natural killer group 2 member D (NKG2D)
on NK cell surfaces [166]. The selenium metabolite known as methylselenol was found to upregulate
two NKG2D ligands on the surface of tumor cells [167]. However, it was not determined if this led
to increased NK cell killing of tumor cells. This feature is important for the detection of tumor cells
by CD8+ T cells since these cells also express NKG2D. In fact, major histocompatibility-I (MHC-I)
present tumor antigens to CD8+ T cells to activate their cytotoxic activities, which is also affected
by methylselenol in cancer cells. In particular, this selenium metabolite was shown to alter redox
metabolism in melanoma cells and lead to increased levels of MHC-I cell surface antigens [168].
This study showed that the actions of methylselenol mimic IFNγ signaling by also upregulating
members of IFNγ responsive genes. However, one must consider the detrimental effects of inducing
oxidative stress in some tissues such as the gut where this can promote tumorigenesis and tumor
progression [85].
Due to the ability to control experimental conditions, rodent models of selenium and cancer have
provided data that may be easier to interpret. However, unless specifically built into the study design,
it is difficult if not impossible to distinguish between the effects of the bioavailable selenium on the
cancer cells themselves versus the immune cells that are either trying to facilitate tumor progression or
trying to eradicate the cancer cells. The mixed results for rodent cancer studies when tumor growth
217
Nutrients 2018, 10, 1203
is the primary endpoint for mouse studies highlight this confounding issue. For example, in our
mouse model study of syngeneic mesothelioma tumors that utilized immune competent animals,
we expected that increasing dietary selenium would hinder tumor progression due to enhanced
anti-cancer immunity. However, the tumors progressed at an accelerated rate in mice that were fed
higher selenium diets due to the pro-reducing capacity in the tumor cells themselves [169]. Other rodent
studies focused on melanoma or breast cancer found different results with higher selenium intake
leading to lower tumor growth concurrent with immune enhancing effects [170]. When immune
responses have been analyzed, the predominant effect is an enhancement of Th1 immunity and a
reduction in regulatory T cells (Tregs) and myeloid-derived suppressor cells that suppress anti-tumor
immunity [171,172]. There are many factors to consider when analyzing the results of the rodent
cancer studies including the type of cancer, strain, and immune status of the rodents, dose and form of
selenium used, and endpoints used for analyses. The generation of new xenograft models as well as
humanized rodents will facilitate these studies when the research field moves forward.
218
Nutrients 2018, 10, 1203
(letters represent the amino acids aspartic acid, histidine, histidine, and cysteine in the catalytic
domain and 6 represents the 6th member of the family) to palmitoylate to the endoplasmic reticulum
(ER) calcium channel protein inositol-1,4,5-trisphosphate receptor (IP3R) [176]. SELENOK itself does
not function as an enzyme but instead binds to DHHC6 to stabilize the acylated intermediate of
this enzyme so that it is not hydrolyzed by water, which does not hydrolyze the thioester bond
between the acyl group and the cysteine residue in DHHC6 before it can transfer the acyl group
to target proteins like IP3R [177]. Other proteins involved in immune functions also depend on
SELENOK/DHHC6 for palmitoylation to carry out their activities including CD36 and Arf-GAP with
SH3 domain, ANK repeat and PH domain-containing protein 2 (ASAP2) [178,179]. Overall, SELENOK
plays an important, non-enzymatic role in regulating immunity by functioning as a cofactor for an
enzyme involved in critical post-translational modifications of proteins.
7. Conclusions
The immune system is one aspect of human health that is impacted by dietary selenium levels and
selenoprotein expression. Under conditions of selenium deficiency, innate and adaptive immune
responses are impaired. The benefits of selenium supplementation to boost immunity against
pathogens, vaccinations, or cancers have been explored and have not provided entirely clear results.
Some of the issues lie in the fact that some pathogens or tumor cells may themselves benefit from higher
levels of selenium. Manipulation of individual selenoproteins may offer a more precise approach
for enhancing the immune system or mitigating chronic inflammation. This approach will require
a comprehensive characterization of the roles for selenoproteins and an unmasking of molecular
mechanisms by which they regulate immune cell functions.
References
1. Franke, K.W. A New Toxicant Occurring Naturally in Certain Samples of Plant Foodstuffs: I. Results
Obtained in Preliminary Feeding Trials: Eight Figures. J. Nutr. 1934, 8, 597–608. [CrossRef]
2. Schwarz, K.; Foltz, C.M. Selenium as an integral part of factor 3 against dietary necrosis liver degeneration.
J. Am. Chem. Soc. 1957, 79, 3292–3293. [CrossRef]
3. Roman, M.; Jitaru, P.; Barbante, C. Selenium biochemistry and its role for human health. Metallomics 2014, 6,
25–54. [CrossRef] [PubMed]
4. Rayman, M.P. Selenium and human health. Lancet 2012, 379, 1256–1268. [CrossRef]
5. Chun, O.K.; Floegel, A.; Chung, S.J.; Chung, C.E.; Song, W.O.; Koo, S.I. Estimation of antioxidant intakes
from diet and supplements in U.S. adults. J. Nutr. 2010, 140, 317–324. [CrossRef] [PubMed]
6. Kipp, A.P.; Strohm, D.; Brigelius-Flohe, R.; Schomburg, L.; Bechthold, A.; Leschik-Bonnet, E.; Heseker, H.;
German Nutrition Society (DGE). Revised reference values for selenium intake. J. Trace Elem. Med. Biol. 2015,
32, 195–199. [CrossRef] [PubMed]
7. Institute of Medicine, Food and Nutrition Board Staff. Vitamin C, Vitamin E, Selenium, and Carotenoids;
National Academy Press: Washington, DC, USA, 2000.
8. Dietary reference values for food energy and nutrients for the United Kingdom. Report of the Panel on
Dietary Reference Values of the Committee on Medical Aspects of Food Policy. Rep. Health Soc. Subj. (Lond.)
1991, 41, 1–210.
9. Sunde, R.A. Selenium. In Modern Nutrition in Health and Disease, 11th ed.; Ross, A.C., Caballero, B.,
Cousins, R.J., Tucker, K.L., Ziegler, T.R., Eds.; Lippincott Williams & Wilkins: Philadelphia, PA, USA,
2012; pp. 225–237.
10. Ashton, K.; Hooper, L.; Harvey, L.J.; Hurst, R.; Casgrain, A.; Fairweather-Tait, S.J. Methods of assessment
of selenium status in humans: A systematic review. Am. J. Clin. Nutr. 2009, 89, 2025S–2039S. [CrossRef]
[PubMed]
219
Nutrients 2018, 10, 1203
11. Combs, G.F., Jr. Biomarkers of selenium status. Nutrients 2015, 7, 2209–2236. [CrossRef] [PubMed]
12. Stoffaneller, R.; Morse, N.L. A review of dietary selenium intake and selenium status in Europe and the
Middle East. Nutrients 2015, 7, 1494–1537. [CrossRef] [PubMed]
13. Finley, J.W. Bioavailability of selenium from foods. Nutr. Rev. 2006, 64, 146–151. [CrossRef] [PubMed]
14. Kassam, S.; Goenaga-Infante, H.; Maharaj, L.; Hiley, C.T.; Juliger, S.; Joel, S.P. Methylseleninic acid inhibits
HDAC activity in diffuse large B-cell lymphoma cell lines. Cancer Chemother. Pharmacol. 2011, 68, 815–821.
[CrossRef] [PubMed]
15. Bera, S.; De Rosa, V.; Rachidi, W.; Diamond, A.M. Does a role for selenium in DNA damage repair explain
apparent controversies in its use in chemoprevention? Mutagenesis 2013, 28, 127–134. [CrossRef] [PubMed]
16. Ip, C.; Thompson, H.J.; Zhu, Z.; Ganther, H.E. In vitro and in vivo studies of methylseleninic acid: Evidence
that a monomethylated selenium metabolite is critical for cancer chemoprevention. Cancer Res. 2000, 60,
2882–2886. [PubMed]
17. Reeves, M.A.; Hoffmann, P.R. The human selenoproteome: Recent insights into functions and regulation.
Cell. Mol. Life Sci. 2009, 66, 2457–2478. [CrossRef] [PubMed]
18. Seyedali, A.; Berry, M.J. Nonsense-mediated decay factors are involved in the regulation of selenoprotein
mRNA levels during selenium deficiency. RNA 2014, 20, 1248–1256. [CrossRef] [PubMed]
19. Lin, H.C.; Yeh, C.W.; Chen, Y.F.; Lee, T.T.; Hsieh, P.Y.; Rusnac, D.V.; Lin, S.Y.; Elledge, S.J.; Zheng, N.; Yen, H.S.
C-Terminal End-Directed Protein Elimination by CRL2 Ubiquitin Ligases. Mol. Cell. 2018, 70, 602–613 e603.
[CrossRef] [PubMed]
20. Burk, R.F.; Hill, K.E. Regulation of Selenium Metabolism and Transport. Annu. Rev. Nutr. 2015, 35, 109–134.
[CrossRef] [PubMed]
21. Kryukov, G.V.; Castellano, S.; Novoselov, S.V.; Lobanov, A.V.; Zehtab, O.; Guigo, R.; Gladyshev, V.N.
Characterization of mammalian selenoproteomes. Science 2003, 300, 1439–1443. [CrossRef] [PubMed]
22. Lubos, E.; Loscalzo, J.; Handy, D.E. Glutathione peroxidase-1 in health and disease: From molecular
mechanisms to therapeutic opportunities. Antioxid. Redox Signal. 2011, 15, 1957–1997. [CrossRef] [PubMed]
23. Lei, X.G.; Cheng, W.H.; McClung, J.P. Metabolic regulation and function of glutathione peroxidase-1.
Annu. Rev. Nutr. 2007, 27, 41–61. [CrossRef] [PubMed]
24. Brigelius-Flohe, R.; Kipp, A. Glutathione peroxidases in different stages of carcinogenesis.
Biochim. Biophys. Acta 2009, 1790, 1555–1568. [CrossRef] [PubMed]
25. Wingler, K.; Brigelius-Flohe, R. Gastrointestinal glutathione peroxidase. Biofactors 1999, 10, 245–249.
[CrossRef] [PubMed]
26. Koyama, H.; Omura, K.; Ejima, A.; Kasanuma, Y.; Watanabe, C.; Satoh, H. Separation of selenium-containing
proteins in human and mouse plasma using tandem high-performance liquid chromatography columns
coupled with inductively coupled plasma-mass spectrometry. Anal. Biochem. 1999, 267, 84–91. [CrossRef]
[PubMed]
27. Chu, F.F.; Esworthy, R.S.; Doroshow, J.H.; Doan, K.; Liu, X.F. Expression of plasma glutathione peroxidase in
human liver in addition to kidney, heart, lung, and breast in humans and rodents. Blood 1992, 79, 3233–3238.
[PubMed]
28. Conrad, M.; Schneider, M.; Seiler, A.; Bornkamm, G.W. Physiological role of phospholipid hydroperoxide
glutathione peroxidase in mammals. Biol. Chem. 2007, 388, 1019–1025. [CrossRef] [PubMed]
29. Ingold, I.; Berndt, C.; Schmitt, S.; Doll, S.; Poschmann, G.; Buday, K.; Roveri, A.; Peng, X.; Porto Freitas, F.;
Seibt, T.; et al. Selenium Utilization by GPX4 Is Required to Prevent Hydroperoxide-Induced Ferroptosis.
Cell 2018, 172, 409–422 e421. [CrossRef] [PubMed]
30. Brigelius-Flohe, R. Glutathione peroxidases and redox-regulated transcription factors. Biol. Chem. 2006, 387,
1329–1335. [CrossRef] [PubMed]
31. Crosley, L.K.; Meplan, C.; Nicol, F.; Rundlof, A.K.; Arner, E.S.; Hesketh, J.E.; Arthur, J.R. Differential
regulation of expression of cytosolic and mitochondrial thioredoxin reductase in rat liver and kidney.
Arch. Biochem. Biophys. 2007, 459, 178–188. [CrossRef] [PubMed]
32. Conrad, M.; Jakupoglu, C.; Moreno, S.G.; Lippl, S.; Banjac, A.; Schneider, M.; Beck, H.; Hatzopoulos, A.K.;
Just, U.; Sinowatz, F.; et al. Essential role for mitochondrial thioredoxin reductase in hematopoiesis, heart
development, and heart function. Mol. Cell. Biol. 2004, 24, 9414–9423. [CrossRef] [PubMed]
220
Nutrients 2018, 10, 1203
33. Su, D.; Novoselov, S.V.; Sun, Q.A.; Moustafa, M.E.; Zhou, Y.; Oko, R.; Hatfield, D.L.; Gladyshev, V.N.
Mammalian selenoprotein thioredoxin-glutathione reductase. Roles in disulfide bond formation and sperm
maturation. J. Biol. Chem. 2005, 280, 26491–26498. [CrossRef] [PubMed]
34. Darras, V.M.; Van Herck, S.L. Iodothyronine deiodinase structure and function: From ascidians to humans.
J. Endocrinol. 2012, 215, 189–206. [CrossRef] [PubMed]
35. Lee, B.C.; Lee, S.G.; Choo, M.K.; Kim, J.H.; Lee, H.M.; Kim, S.; Fomenko, D.E.; Kim, H.Y.; Park, J.M.;
Gladyshev, V.N. Selenoprotein MsrB1 promotes anti-inflammatory cytokine gene expression in macrophages
and controls immune response in vivo. Sci. Rep. 2017, 7, 5119. [CrossRef] [PubMed]
36. Fomenko, D.E.; Novoselov, S.V.; Natarajan, S.K.; Lee, B.C.; Koc, A.; Carlson, B.A.; Lee, T.H.; Kim, H.Y.;
Hatfield, D.L.; Gladyshev, V.N. MsrB1 (methionine-R-sulfoxide reductase 1) knock-out mice: Roles of MsrB1
in redox regulation and identification of a novel selenoprotein form. J. Biol. Chem. 2009, 284, 5986–5993.
[CrossRef] [PubMed]
37. Labunskyy, V.M.; Hatfield, D.L.; Gladyshev, V.N. The Sep15 protein family: Roles in disulfide bond formation
and quality control in the endoplasmic reticulum. IUBMB Life 2007, 59, 1–5. [CrossRef] [PubMed]
38. Yim, S.H.; Everley, R.A.; Schildberg, F.A.; Lee, S.G.; Orsi, A.; Barbati, Z.R.; Karatepe, K.; Fomenko, D.E.;
Tsuji, P.A.; Luo, H.R.; et al. Role of Selenof as a Gatekeeper of Secreted Disulfide-Rich Glycoproteins. Cell. Rep.
2018, 23, 1387–1398. [CrossRef] [PubMed]
39. Panee, J.; Stoytcheva, Z.R.; Liu, W.; Berry, M.J. Selenoprotein H is a redox-sensing high mobility group family
DNA-binding protein that up-regulates genes involved in glutathione synthesis and phase II detoxification.
J. Biol. Chem. 2007, 282, 23759–23765. [CrossRef] [PubMed]
40. Novoselov, S.V.; Kryukov, G.V.; Xu, X.M.; Carlson, B.A.; Hatfield, D.L.; Gladyshev, V.N. Selenoprotein H is a
nucleolar thioredoxin-like protein with a unique expression pattern. J. Biol. Chem. 2007, 282, 11960–11968.
[CrossRef] [PubMed]
41. Horibata, Y.; Hirabayashi, Y. Identification and characterization of human ethanolaminephosphotransferase1.
J. Lipid Res. 2007, 48, 503–508. [CrossRef] [PubMed]
42. Verma, S.; Hoffmann, F.W.; Kumar, M.; Huang, Z.; Roe, K.; Nguyen-Wu, E.; Hashimoto, A.S.; Hoffmann, P.R.
Selenoprotein K knockout mice exhibit deficient calcium flux in immune cells and impaired immune
responses. J. Immunol. 2011, 186, 2127–2137. [CrossRef] [PubMed]
43. Fredericks, G.J.; Hoffmann, P.R. Selenoprotein K and protein palmitoylation. Antioxid. Redox Signal. 2015, 23,
854–862. [CrossRef] [PubMed]
44. Pitts, M.W.; Reeves, M.A.; Hashimoto, A.C.; Ogawa, A.; Kremer, P.; Seale, L.A.; Berry, M.J. Deletion of
selenoprotein M leads to obesity without cognitive deficits. J. Biol. Chem. 2013, 288, 26121–26134. [CrossRef]
[PubMed]
45. Lescure, A.; Rederstorff, M.; Krol, A.; Guicheney, P.; Allamand, V. Selenoprotein function and muscle disease.
Biochim. Biophys. Acta 2009, 1790, 1569–1574. [CrossRef] [PubMed]
46. Castets, P.; Lescure, A.; Guicheney, P.; Allamand, V. Selenoprotein N in skeletal muscle: From diseases to
function. J. Mol. Med. (Berl.) 2012, 90, 1095–1107. [CrossRef] [PubMed]
47. Han, S.J.; Lee, B.C.; Yim, S.H.; Gladyshev, V.N.; Lee, S.R. Characterization of mammalian selenoprotein o:
A redox-active mitochondrial protein. PLoS ONE 2014, 9, e95518. [CrossRef] [PubMed]
48. Burk, R.F.; Hill, K.E. Selenoprotein P-expression, functions, and roles in mammals. Biochim. Biophys. Acta
2009, 1790, 1441–1447. [CrossRef] [PubMed]
49. Ye, Y.; Shibata, Y.; Yun, C.; Ron, D.; Rapoport, T.A. A membrane protein complex mediates retro-translocation
from the ER lumen into the cytosol. Nature 2004, 429, 841–847. [CrossRef] [PubMed]
50. Turanov, A.A.; Shchedrina, V.A.; Everley, R.A.; Lobanov, A.V.; Yim, S.H.; Marino, S.M.; Gygi, S.P.;
Hatfield, D.L.; Gladyshev, V.N. Selenoprotein S is involved in maintenance and transport of multiprotein
complexes. Biochem. J. 2014, 462, 555–565. [CrossRef] [PubMed]
51. Boukhzar, L.; Hamieh, A.; Cartier, D.; Tanguy, Y.; Alsharif, I.; Castex, M.; Arabo, A.; El Hajji, S.; Bonnet, J.J.;
Errami, M.; et al. Selenoprotein T Exerts an Essential Oxidoreductase Activity That Protects Dopaminergic
Neurons in Mouse Models of Parkinson’s Disease. Antioxid. Redox Signal. 2016, 24, 557–574. [CrossRef]
[PubMed]
52. Jeon, Y.H.; Park, Y.H.; Lee, J.H.; Hong, J.H.; Kim, I.Y. Selenoprotein W enhances skeletal muscle differentiation
by inhibiting TAZ binding to 14-3-3 protein. Biochim. Biophys. Acta 2014, 1843, 1356–1364. [CrossRef]
[PubMed]
221
Nutrients 2018, 10, 1203
53. Xu, X.M.; Carlson, B.A.; Irons, R.; Mix, H.; Zhong, N.; Gladyshev, V.N.; Hatfield, D.L. Selenophosphate
synthetase 2 is essential for selenoprotein biosynthesis. Biochem. J. 2007, 404, 115–120. [CrossRef] [PubMed]
54. Bosl, M.R.; Takaku, K.; Oshima, M.; Nishimura, S.; Taketo, M.M. Early embryonic lethality caused by targeted
disruption of the mouse selenocysteine tRNA gene (Trsp). Proc. Natl. Acad. Sci. USA 1997, 94, 5531–5534.
[CrossRef] [PubMed]
55. Yant, L.J.; Ran, Q.; Rao, L.; Van Remmen, H.; Shibatani, T.; Belter, J.G.; Motta, L.; Richardson, A.; Prolla, T.A.
The selenoprotein GPX4 is essential for mouse development and protects from radiation and oxidative
damage insults. Free Radic. Biol. Med. 2003, 34, 496–502. [CrossRef]
56. Jakupoglu, C.; Przemeck, G.K.; Schneider, M.; Moreno, S.G.; Mayr, N.; Hatzopoulos, A.K.; de Angelis, M.H.;
Wurst, W.; Bornkamm, G.W.; Brielmeier, M.; et al. Cytoplasmic thioredoxin reductase is essential for
embryogenesis but dispensable for cardiac development. Mol. Cell. Biol. 2005, 25, 1980–1988. [CrossRef]
[PubMed]
57. Ferreiro, A.; Quijano-Roy, S.; Pichereau, C.; Moghadaszadeh, B.; Goemans, N.; Bonnemann, C.; Jungbluth, H.;
Straub, V.; Villanova, M.; Leroy, J.P.; et al. Mutations of the selenoprotein N gene, which is implicated in
rigid spine muscular dystrophy, cause the classical phenotype of multiminicore disease: Reassessing the
nosology of early-onset myopathies. Am. J. Hum. Genet. 2002, 71, 739–749. [CrossRef] [PubMed]
58. Moghadaszadeh, B.; Petit, N.; Jaillard, C.; Brockington, M.; Quijano Roy, S.; Merlini, L.; Romero, N.;
Estournet, B.; Desguerre, I.; Chaigne, D.; et al. Mutations in SEPN1 cause congenital muscular dystrophy
with spinal rigidity and restrictive respiratory syndrome. Nat. Genet. 2001, 29, 17–18. [CrossRef] [PubMed]
59. Jurynec, M.J.; Xia, R.; Mackrill, J.J.; Gunther, D.; Crawford, T.; Flanigan, K.M.; Abramson, J.J.; Howard, M.T.;
Grunwald, D.J. Selenoprotein N is required for ryanodine receptor calcium release channel activity in human
and zebrafish muscle. Proc. Natl. Acad. Sci. USA 2008, 105, 12485–12490. [CrossRef] [PubMed]
60. Hornberger, T.A.; McLoughlin, T.J.; Leszczynski, J.K.; Armstrong, D.D.; Jameson, R.R.; Bowen, P.E.;
Hwang, E.S.; Hou, H.; Moustafa, M.E.; Carlson, B.A.; et al. Selenoprotein-deficient transgenic mice exhibit
enhanced exercise-induced muscle growth. J. Nutr. 2003, 133, 3091–3097. [CrossRef] [PubMed]
61. Hernandez, A.; St Germain, D.L. Thyroid hormone deiodinases: Physiology and clinical disorders.
Curr. Opin. Pediatr. 2003, 15, 416–420. [CrossRef] [PubMed]
62. Kato, T.; Read, R.; Rozga, J.; Burk, R.F. Evidence for intestinal release of absorbed selenium in a form with
high hepatic extraction. Am. J. Physiol. 1992, 262, G854–G858. [CrossRef] [PubMed]
63. Hill, K.E.; Zhou, J.; McMahan, W.J.; Motley, A.K.; Atkins, J.F.; Gesteland, R.F.; Burk, R.F. Deletion of
selenoprotein P alters distribution of selenium in the mouse. J. Biol. Chem. 2003, 278, 13640–13646. [CrossRef]
[PubMed]
64. Schomburg, L.; Schweizer, U.; Holtmann, B.; Flohe, L.; Sendtner, M.; Kohrle, J. Gene disruption discloses role
of selenoprotein P in selenium delivery to target tissues. Biochem. J. 2003, 370, 397–402. [CrossRef] [PubMed]
65. Behne, D.; Hilmert, H.; Scheid, S.; Gessner, H.; Elger, W. Evidence for specific selenium target tissues and
new biologically important selenoproteins. Biochim. Biophys. Acta 1988, 966, 12–21. [CrossRef]
66. Valentine, W.M.; Hill, K.E.; Austin, L.M.; Valentine, H.L.; Goldowitz, D.; Burk, R.F. Brainstem axonal
degeneration in mice with deletion of selenoprotein p. Toxicol. Pathol. 2005, 33, 570–576. [CrossRef]
[PubMed]
67. Hill, K.E.; Zhou, J.; McMahan, W.J.; Motley, A.K.; Burk, R.F. Neurological dysfunction occurs in mice with
targeted deletion of the selenoprotein P gene. J. Nutr. 2004, 134, 157–161. [CrossRef] [PubMed]
68. Schweizer, U.; Michaelis, M.; Kohrle, J.; Schomburg, L. Efficient selenium transfer from mother to offspring
in selenoprotein-P-deficient mice enables dose-dependent rescue of phenotypes associated with selenium
deficiency. Biochem. J. 2004, 378, 21–26. [CrossRef] [PubMed]
69. Nonn, L.; Williams, R.R.; Erickson, R.P.; Powis, G. The absence of mitochondrial thioredoxin 2 causes massive
apoptosis, exencephaly, and early embryonic lethality in homozygous mice. Mol. Cell. Biol. 2003, 23, 916–922.
[CrossRef] [PubMed]
70. Flohe, L.; Gunzler, W.A.; Schock, H.H. Glutathione peroxidase: A selenoenzyme. FEBS Lett. 1973, 32, 132–134.
[CrossRef]
71. Rotruck, J.T.; Pope, A.L.; Ganther, H.E.; Swanson, A.B.; Hafeman, D.G.; Hoekstra, W.G. Selenium:
Biochemical role as a component of glutathione peroxidase. Science 1973, 179, 588–590. [CrossRef] [PubMed]
222
Nutrients 2018, 10, 1203
72. Conrad, M. Transgenic mouse models for the vital selenoenzymes cytosolic thioredoxin reductase,
mitochondrial thioredoxin reductase and glutathione peroxidase 4. Biochim. Biophys. Acta 2009, 1790,
1575–1585. [CrossRef] [PubMed]
73. Schomburg, L.; Schweizer, U. Hierarchical regulation of selenoprotein expression and sex-specific effects of
selenium. Biochim. Biophys. Acta 2009, 1790, 1453–1462. [CrossRef] [PubMed]
74. Burk, R.F.; Hill, K.E.; Awad, J.A.; Morrow, J.D.; Lyons, P.R. Liver and kidney necrosis in selenium-deficient
rats depleted of glutathione. Lab. Investig. 1995, 72, 723–730. [PubMed]
75. Ge, K.; Xue, A.; Bai, J.; Wang, S. Keshan disease-an endemic cardiomyopathy in China. Virchows Arch. A
Pathol. Anat. Histopathol. 1983, 401, 1–15. [CrossRef] [PubMed]
76. Beck, M.A.; Levander, O.A.; Handy, J. Selenium deficiency and viral infection. J. Nutr. 2003, 133, 1463S–1467S.
[CrossRef] [PubMed]
77. Sokoloff, L. Acquired chondronecrosis. Ann. Rheum. Dis. 1990, 49, 262–264. [CrossRef] [PubMed]
78. Wang, S.J.; Guo, X.; Zuo, H.; Zhang, Y.G.; Xu, P.; Ping, Z.G.; Zhang, Z.; Geng, D. Chondrocyte apoptosis
and expression of Bcl-2, Bax, Fas, and iNOS in articular cartilage in patients with Kashin-Beck disease.
J. Rheumatol. 2006, 33, 615–619. [PubMed]
79. Wang, S.J.; Guo, X.; Ren, F.L.; Zhang, Y.G.; Zhang, Z.T.; Zhang, F.J.; Geng, D. Comparison of apoptosis of
articular chondrocytes in the pathogenesis of Kashin-beck disease and primary osteoarthritis. Zhongguo Yi
Xue Ke Xue Yuan Xue Bao 2006, 28, 267–270. (In Chinese) [PubMed]
80. Wang, Y.; Guo, X.; Zhang, Z.T.; Wang, M.; Wang, S.J. Expression of Caspase-8 and Bcl-2 in the cartilage loose
bodies in patients with Kashin-Beck disease. Nan Fang Yi Ke Da Xue Xue Bao 2011, 31, 1314–1317. (In Chinese)
[PubMed]
81. Yao, Y.; Pei, F.; Kang, P. Selenium, iodine, and the relation with Kashin-Beck disease. Nutrition 2011, 27,
1095–1100. [CrossRef] [PubMed]
82. Moreno-Reyes, R.; Suetens, C.; Mathieu, F.; Begaux, F.; Zhu, D.; Rivera, M.T.; Boelaert, M.; Neve, J.;
Perlmutter, N.; Vanderpas, J. Kashin-Beck osteoarthropathy in rural Tibet in relation to selenium and
iodine status. N. Engl. J. Med. 1998, 339, 1112–1120. [CrossRef] [PubMed]
83. Barrett, C.W.; Reddy, V.K.; Short, S.P.; Motley, A.K.; Lintel, M.K.; Bradley, A.M.; Freeman, T.; Vallance, J.;
Ning, W.; Parang, B.; et al. Selenoprotein P influences colitis-induced tumorigenesis by mediating stemness
and oxidative damage. J. Clin. Investig. 2015, 125, 2646–2660. [CrossRef] [PubMed]
84. Hamid, M.; Abdulrahim, Y.; Liu, D.; Qian, G.; Khan, A.; Huang, K. The Hepatoprotective Effect of
Selenium-Enriched Yeast and Gum Arabic Combination on Carbon Tetrachloride-Induced Chronic Liver
Injury in Rats. J. Food Sci. 2018, 83, 525–534. [CrossRef] [PubMed]
85. Barrett, C.W.; Short, S.P.; Williams, C.S. Selenoproteins and oxidative stress-induced inflammatory
tumorigenesis in the gut. Cell. Mol. Life Sci. 2017, 74, 607–616. [CrossRef] [PubMed]
86. Nettleford, S.K.; Prabhu, K.S. Selenium and Selenoproteins in Gut Inflammation-A Review.
Antioxidants (Basel) 2018, 7, 36. [CrossRef] [PubMed]
87. Zhang, Z.; Gao, X.; Cao, Y.; Jiang, H.; Wang, T.; Song, X.; Guo, M.; Zhang, N. Selenium Deficiency Facilitates
Inflammation Through the Regulation of TLR4 and TLR4-Related Signaling Pathways in the Mice Uterus.
Inflammation 2015, 38, 1347–1356. [CrossRef] [PubMed]
88. Gao, X.; Zhang, Z.; Li, Y.; Shen, P.; Hu, X.; Cao, Y.; Zhang, N. Selenium Deficiency Facilitates Inflammation
Following S. aureus Infection by Regulating TLR2-Related Pathways in the Mouse Mammary Gland.
Biol. Trace Elem. Res. 2016, 172, 449–457. [CrossRef] [PubMed]
89. Hoffmann, P.R.; Jourdan-Le Saux, C.; Hoffmann, F.W.; Chang, P.S.; Bollt, O.; He, Q.; Tam, E.K.; Berry, M.J.
A role for dietary selenium and selenoproteins in allergic airway inflammation. J. Immunol. 2007, 179,
3258–3267. [CrossRef] [PubMed]
90. Tsuji, P.A.; Carlson, B.A.; Anderson, C.B.; Seifried, H.E.; Hatfield, D.L.; Howard, M.T. Dietary Selenium
Levels Affect Selenoprotein Expression and Support the Interferon-gamma and IL-6 Immune Response
Pathways in Mice. Nutrients 2015, 7, 6529–6549. [CrossRef] [PubMed]
91. Huang, Z.; Rose, A.H.; Hoffmann, P.R. The role of selenium in inflammation and immunity: From molecular
mechanisms to therapeutic opportunities. Antioxid. Redox Signal. 2012, 16, 705–743. [CrossRef] [PubMed]
223
Nutrients 2018, 10, 1203
92. Bentley-Hewitt, K.L.; Chen, R.K.; Lill, R.E.; Hedderley, D.I.; Herath, T.D.; Matich, A.J.; McKenzie, M.J.
Consumption of selenium-enriched broccoli increases cytokine production in human peripheral blood
mononuclear cells stimulated ex vivo, a preliminary human intervention study. Mol. Nutr. Food Res. 2014, 58,
2350–2357. [CrossRef] [PubMed]
93. Broome, C.S.; McArdle, F.; Kyle, J.A.; Andrews, F.; Lowe, N.M.; Hart, C.A.; Arthur, J.R.; Jackson, M.J.
An increase in selenium intake improves immune function and poliovirus handling in adults with marginal
selenium status. Am. J. Clin. Nutr. 2004, 80, 154–162. [CrossRef] [PubMed]
94. Ivory, K.; Prieto, E.; Spinks, C.; Armah, C.N.; Goldson, A.J.; Dainty, J.R.; Nicoletti, C. Selenium
supplementation has beneficial and detrimental effects on immunity to influenza vaccine in older adults.
Clin. Nutr. 2017, 36, 407–415. [CrossRef] [PubMed]
95. Meplan, C.; Johnson, I.T.; Polley, A.C.; Cockell, S.; Bradburn, D.M.; Commane, D.M.; Arasaradnam, R.P.;
Mulholland, F.; Zupanic, A.; Mathers, J.C.; et al. Transcriptomics and proteomics show that selenium affects
inflammation, cytoskeleton, and cancer pathways in human rectal biopsies. FASEB J. 2016, 30, 2812–2825.
[CrossRef] [PubMed]
96. Mahmoodpoor, A.; Hamishehkar, H.; Shadvar, K.; Ostadi, Z.; Sanaie, S.; Saghaleini, S.H.; Nader, N.D.
The Effect of Intravenous Selenium on Oxidative Stress in Critically Ill Patients with Acute Respiratory
Distress Syndrome. Immunol. Investig. 2018, 1–13. [CrossRef] [PubMed]
97. Geisberger, R.; Kiermayer, C.; Homig, C.; Conrad, M.; Schmidt, J.; Zimber-Strobl, U.; Brielmeier, M.
B- and T-cell-specific inactivation of thioredoxin reductase 2 does not impair lymphocyte development
and maintenance. Biol. Chem. 2007, 388, 1083–1090. [CrossRef] [PubMed]
98. Shrimali, R.K.; Irons, R.D.; Carlson, B.A.; Sano, Y.; Gladyshev, V.N.; Park, J.M.; Hatfield, D.L. Selenoproteins
mediate T cell immunity through an antioxidant mechanism. J. Biol. Chem. 2008, 283, 20181–20185. [CrossRef]
[PubMed]
99. Wichman, J.; Winther, K.H.; Bonnema, S.J.; Hegedus, L. Selenium Supplementation Significantly Reduces
Thyroid Autoantibody Levels in Patients with Chronic Autoimmune Thyroiditis: A Systematic Review and
Meta-Analysis. Thyroid 2016, 26, 1681–1692. [CrossRef] [PubMed]
100. McLachlan, S.M.; Aliesky, H.; Banuelos, B.; Hee, S.S.Q.; Rapoport, B. Variable Effects of Dietary Selenium
in Mice That Spontaneously Develop a Spectrum of Thyroid Autoantibodies. Endocrinology 2017, 158,
3754–3764. [CrossRef] [PubMed]
101. Hoffmann, F.W.; Hashimoto, A.C.; Shafer, L.A.; Dow, S.; Berry, M.J.; Hoffmann, P.R. Dietary selenium
modulates activation and differentiation of CD4+ T cells in mice through a mechanism involving cellular
free thiols. J. Nutr. 2010, 140, 1155–1161. [CrossRef] [PubMed]
102. Mahdavi, M.; Mavandadnejad, F.; Yazdi, M.H.; Faghfuri, E.; Hashemi, H.; Homayouni-Oreh, S.; Farhoudi, R.;
Shahverdi, A.R. Oral administration of synthetic selenium nanoparticles induced robust Th1 cytokine pattern
after HBs antigen vaccination in mouse model. J. Infect. Public Health 2017, 10, 102–109. [CrossRef] [PubMed]
103. Roy, M.; Kiremidjian-Schumacher, L.; Wishe, H.I.; Cohen, M.W.; Stotzky, G. Supplementation with selenium
restores age-related decline in immune cell function. Proc. Soc. Exp. Biol. Med. 1995, 209, 369–375. [CrossRef]
[PubMed]
104. Hawkes, W.C.; Kelley, D.S.; Taylor, P.C. The effects of dietary selenium on the immune system in healthy
men. Biol. Trace Elem. Res. 2001, 81, 189–213. [CrossRef]
105. Nelson, S.M.; Lei, X.; Prabhu, K.S. Selenium levels affect the IL-4-induced expression of alternative activation
markers in murine macrophages. J. Nutr. 2011, 141, 1754–1761. [CrossRef] [PubMed]
106. Nelson, S.M.; Shay, A.E.; James, J.L.; Carlson, B.A.; Urban, J.F., Jr.; Prabhu, K.S. Selenoprotein Expression
in Macrophages Is Critical for Optimal Clearance of Parasitic Helminth Nippostrongylus brasiliensis.
J. Biol. Chem. 2016, 291, 2787–2798. [CrossRef] [PubMed]
107. Carlson, B.A.; Yoo, M.H.; Shrimali, R.K.; Irons, R.; Gladyshev, V.N.; Hatfield, D.L.; Park, J.M. Role of
selenium-containing proteins in T-cell and macrophage function. Proc. Nutr. Soc. 2010, 69, 300–310.
[CrossRef] [PubMed]
108. Safir, N.; Wendel, A.; Saile, R.; Chabraoui, L. The effect of selenium on immune functions of J774.1 cells.
Clin. Chem. Lab. Med. 2003, 41, 1005–1011. [CrossRef] [PubMed]
109. Aribi, M.; Meziane, W.; Habi, S.; Boulatika, Y.; Marchandin, H.; Aymeric, J.L. Macrophage Bactericidal
Activities against Staphylococcus aureus Are Enhanced In Vivo by Selenium Supplementation in a
Dose-Dependent Manner. PLoS ONE 2015, 10, e0135515. [CrossRef] [PubMed]
224
Nutrients 2018, 10, 1203
110. Bi, C.L.; Wang, H.; Wang, Y.J.; Sun, J.; Dong, J.S.; Meng, X.; Li, J.J. Selenium inhibits Staphylococcus
aureus-induced inflammation by suppressing the activation of the NF-kappaB and MAPK signalling
pathways in RAW264.7 macrophages. Eur. J. Pharmacol. 2016, 780, 159–165. [CrossRef] [PubMed]
111. Kose, S.A.; Naziroglu, M. Selenium reduces oxidative stress and calcium entry through TRPV1 channels
in the neutrophils of patients with polycystic ovary syndrome. Biol. Trace Elem. Res. 2014, 158, 136–142.
[CrossRef] [PubMed]
112. Ravaglia, G.; Forti, P.; Maioli, F.; Bastagli, L.; Facchini, A.; Mariani, E.; Savarino, L.; Sassi, S.; Cucinotta, D.;
Lenaz, G. Effect of micronutrient status on natural killer cell immune function in healthy free-living subjects
aged >/=90 y. Am. J. Clin. Nutr. 2000, 71, 590–598. [CrossRef] [PubMed]
113. Kiremidjian-Schumacher, L.; Roy, M.; Wishe, H.I.; Cohen, M.W.; Stotzky, G. Supplementation with selenium
augments the functions of natural killer and lymphokine-activated killer cells. Biol. Trace Elem. Res. 1996, 52,
227–239. [CrossRef] [PubMed]
114. Enqvist, M.; Nilsonne, G.; Hammarfjord, O.; Wallin, R.P.; Bjorkstrom, N.K.; Bjornstedt, M.; Hjerpe, A.;
Ljunggren, H.G.; Dobra, K.; Malmberg, K.J.; et al. Selenite induces posttranscriptional blockade of HLA-E
expression and sensitizes tumor cells to CD94/NKG2A-positive NK cells. J. Immunol. 2011, 187, 3546–3554.
[CrossRef] [PubMed]
115. Alvarado, C.; Alvarez, P.; Jimenez, L.; De la Fuente, M. Improvement of leukocyte functions in young
prematurely aging mice after a 5-week ingestion of a diet supplemented with biscuits enriched in antioxidants.
Antioxid. Redox Signal. 2005, 7, 1203–1210. [CrossRef] [PubMed]
116. Wang, C.; Wang, H.; Luo, J.; Hu, Y.; Wei, L.; Duan, M.; He, H. Selenium deficiency impairs host innate
immune response and induces susceptibility to Listeria monocytogenes infection. BMC Immunol. 2009, 10,
55. [CrossRef] [PubMed]
117. De Freitas, M.R.B.; da Costa, C.M.B.; Pereira, L.M.; do Prado, J.C.J.; Sala, M.A.; Abrahao, A.A.C. The treatment
with selenium increases placental parasitismin pregnant Wistar rats infected with the Y strain of Trypanosoma
cruzi. Immunobiology 2018. [CrossRef] [PubMed]
118. Nelson, S.M.; Shay, A.E.; James, J.L.; Carlson, B.A.; Urban, J.F., Jr.; Prabhu, K.S. Selenoprotein Expression
in Macrophages Is Critical for Optimal Clearance of Parasitic Helminth Nippostrongylus brasiliensis.
J. Biol. Chem. 2013, 291, 2787–2798. [CrossRef]
119. Smith, A.D.; Cheung, L.; Beshah, E.; Shea-Donohue, T.; Urban, J.F., Jr. Selenium status alters the immune
response and expulsion of adult Heligmosomoides bakeri worms in mice. Infect. Immun. 2013, 81, 2546–2553.
[CrossRef] [PubMed]
120. Wiehe, L.; Cremer, M.; Wisniewska, M.; Becker, N.P.; Rijntjes, E.; Martitz, J.; Hybsier, S.; Renko, K.; Buhrer, C.;
Schomburg, L. Selenium status in neonates with connatal infection. Br. J. Nutr. 2016, 116, 504–513. [CrossRef]
[PubMed]
121. Liu, Y.; Qiu, C.; Li, W.; Mu, W.; Li, C.; Guo, M. Selenium Plays a Protective Role in Staphylococcus
aureus-Induced Endometritis in the Uterine Tissue of Rats. Biol. Trace Elem. Res. 2016, 173, 345–353.
[CrossRef] [PubMed]
122. Varsi, K.; Bolann, B.; Torsvik, I.; Rosvold Eik, T.C.; Hol, P.J.; Bjorke-Monsen, A.L. Impact of Maternal Selenium
Status on Infant Outcome during the First 6 Months of Life. Nutrients 2017, 9, 486. [CrossRef] [PubMed]
123. Yoshizawa, S.; Bock, A. The many levels of control on bacterial selenoprotein synthesis. Biochim. Biophys. Acta
2009, 1790, 1404–1414. [CrossRef] [PubMed]
124. Grobler, L.; Nagpal, S.; Sudarsanam, T.D.; Sinclair, D. Nutritional supplements for people being treated for
active tuberculosis. Cochrane Database Syst. Rev. 2016. [CrossRef] [PubMed]
125. Ramakrishnan, K.; Shenbagarathai, R.; Kavitha, K.; Thirumalaikolundusubramanian, P.; Rathinasabapati, R.
Selenium levels in persons with HIV/tuberculosis in India, Madurai City. Clin. Lab. 2012, 58, 165–168.
[PubMed]
126. Eick, F.; Maleta, K.; Govasmark, E.; Duttaroy, A.K.; Bjune, A.G. Food intake of selenium and sulphur amino
acids in tuberculosis patients and healthy adults in Malawi. Int. J. Tuberc. Lung Dis. 2009, 13, 1313–1315.
[PubMed]
127. Seyedrezazadeh, E.; Ostadrahimi, A.; Mahboob, S.; Assadi, Y.; Ghaemmagami, J.; Pourmogaddam, M. Effect
of vitamin E and selenium supplementation on oxidative stress status in pulmonary tuberculosis patients.
Respirology 2008, 13, 294–298. [CrossRef] [PubMed]
225
Nutrients 2018, 10, 1203
128. Sargazi, A.; Gharebagh, R.A.; Sargazi, A.; Aali, H.; Oskoee, H.O.; Sepehri, Z. Role of essential trace elements
in tuberculosis infection: A review article. Indian J. Tuberc. 2017, 64, 246–251. [CrossRef] [PubMed]
129. Jaquess, P.A.; Smalley, D.L.; Duckworth, J.K. Enhanced growth of Mycobacterium tuberculosis in the presence
of selenium. Am. J. Clin. Pathol. 1981, 75, 209–210. [CrossRef] [PubMed]
130. Steinbrenner, H.; Al-Quraishy, S.; Dkhil, M.A.; Wunderlich, F.; Sies, H. Dietary selenium in adjuvant therapy
of viral and bacterial infections. Adv. Nutr. 2015, 6, 73–82. [CrossRef] [PubMed]
131. Puertollano, M.A.; Puertollano, E.; de Cienfuegos, G.A.; de Pablo, M.A. Dietary antioxidants: Immunity and
host defense. Curr. Top. Med. Chem. 2011, 11, 1752–1766. [CrossRef] [PubMed]
132. Ko, W.S.; Guo, C.H.; Yeh, M.S.; Lin, L.Y.; Hsu, G.S.; Chen, P.C.; Luo, M.C.; Lin, C.Y. Blood micronutrient,
oxidative stress, and viral load in patients with chronic hepatitis C. World J. Gastroenterol. 2005, 11, 4697–4702.
[CrossRef] [PubMed]
133. Beck, M.A. Selenium and host defence towards viruses. Proc. Nutr. Soc. 1999, 58, 707–711. [CrossRef]
[PubMed]
134. Jackson, M.J.; Dillon, S.A.; Broome, C.S.; McArdle, A.; Hart, C.A.; McArdle, F. Are there functional
consequences of a reduction in selenium intake in UK subjects? Proc. Nutr. Soc. 2004, 63, 513–517. [CrossRef]
[PubMed]
135. Girodon, F.; Galan, P.; Monget, A.L.; Boutron-Ruault, M.C.; Brunet-Lecomte, P.; Preziosi, P.; Arnaud, J.;
Manuguerra, J.C.; Herchberg, S. Impact of trace elements and vitamin supplementation on immunity and
infections in institutionalized elderly patients: A randomized controlled trial. MIN. VIT. AOX. geriatric
network. Arch. Intern. Med. 1999, 159, 748–754. [CrossRef] [PubMed]
136. Cohen, M.S.; Hellmann, N.; Levy, J.A.; DeCock, K.; Lange, J. The spread, treatment, and prevention of HIV-1:
Evolution of a global pandemic. J. Clin. Investig. 2008, 118, 1244–1254. [CrossRef] [PubMed]
137. Shivakoti, R.; Christian, P.; Yang, W.T.; Gupte, N.; Mwelase, N.; Kanyama, C.; Pillay, S.; Samaneka, W.;
Santos, B.; Poongulali, S.; et al. Prevalence and risk factors of micronutrient deficiencies pre- and
post-antiretroviral therapy (ART) among a diverse multicountry cohort of HIV-infected adults. Clin. Nutr.
2016, 35, 183–189. [CrossRef] [PubMed]
138. Anyabolu, H.C.; Adejuyigbe, E.A.; Adeodu, O.O. Serum Micronutrient Status of Haart-Naive, HIV Infected
Children in South Western Nigeria: A Case Controlled Study. AIDS Res. Treat. 2014, 2014, 351043. [CrossRef]
[PubMed]
139. Shivakoti, R.; Ewald, E.R.; Gupte, N.; Yang, W.T.; Kanyama, C.; Cardoso, S.W.; Santos, B.; Supparatpinyo, K.;
Badal-Faesen, S.; Lama, J.R.; et al. Effect of baseline micronutrient and inflammation status on CD4 recovery
post-cART initiation in the multinational PEARLS trial. Clin. Nutr. 2018. [CrossRef] [PubMed]
140. Dworkin, B.M. Selenium deficiency in HIV infection and the acquired immunodeficiency syndrome (AIDS).
Chem. Biol. Interact. 1994, 91, 181–186. [CrossRef]
141. Stone, C.A.; Kawai, K.; Kupka, R.; Fawzi, W.W. Role of selenium in HIV infection. Nutr. Rev. 2010, 68,
671–681. [CrossRef] [PubMed]
142. Combs, G.F., Jr.; Watts, J.C.; Jackson, M.I.; Johnson, L.K.; Zeng, H.; Scheett, A.J.; Uthus, E.O.; Schomburg, L.;
Hoeg, A.; Hoefig, C.S.; et al. Determinants of selenium status in healthy adults. Nutr. J. 2011, 10, 75.
[CrossRef] [PubMed]
143. Irlam, J.H.; Siegfried, N.; Visser, M.E.; Rollins, N.C. Micronutrient supplementation for children with HIV
infection. Cochrane Database Syst. Rev. 2013. [CrossRef] [PubMed]
144. Hileman, C.O.; Dirajlal-Fargo, S.; Lam, S.K.; Kumar, J.; Lacher, C.; Combs, G.F., Jr.; McComsey, G.A. Plasma
Selenium Concentrations Are Sufficient and Associated with Protease Inhibitor Use in Treated HIV-Infected
Adults. J. Nutr. 2015, 145, 2293–2299. [CrossRef] [PubMed]
145. Akinboro, A.O.; Onayemi, O.; Ayodele, O.E.; Mejiuni, A.D.; Atiba, A.S. The impacts of first line highly active
antiretroviral therapy on serum selenium, CD4 count and body mass index: A cross sectional and short
prospective study. Pan. Afr. Med. J. 2013, 15, 97. [CrossRef] [PubMed]
146. Flax, V.L.; Adair, L.S.; Allen, L.H.; Shahab-Ferdows, S.; Hampel, D.; Chasela, C.S.; Tegha, G.; Daza, E.J.;
Corbett, A.; Davis, N.L.; et al. Plasma Micronutrient Concentrations Are Altered by Antiretroviral Therapy
and Lipid-Based Nutrient Supplements in Lactating HIV-Infected Malawian Women. J. Nutr. 2015, 145,
1950–1957. [PubMed]
147. Baum, M.K.; Shor-Posner, G. Micronutrient status in relationship to mortality in HIV-1 disease. Nutr. Rev.
1998, 56, S135–S139. [CrossRef] [PubMed]
226
Nutrients 2018, 10, 1203
148. Kamwesiga, J.; Mutabazi, V.; Kayumba, J.; Tayari, J.C.; Uwimbabazi, J.C.; Batanage, G.; Uwera, G.;
Baziruwiha, M.; Ntizimira, C.; Murebwayire, A.; et al. Effect of selenium supplementation on CD4+ T-cell
recovery, viral suppression and morbidity of HIV-infected patients in Rwanda: A randomized controlled
trial. AIDS 2015, 29, 1045–1052. [CrossRef] [PubMed]
149. Sappey, C.; Legrand-Poels, S.; Best-Belpomme, M.; Favier, A.; Rentier, B.; Piette, J. Stimulation of glutathione
peroxidase activity decreases HIV type 1 activation after oxidative stress. AIDS Res. Hum. Retroviruses 1994,
10, 1451–1461. [CrossRef] [PubMed]
150. Gupta, S.; Narang, R.; Krishnaswami, K.; Yadav, S. Plasma selenium level in cancer patients. Indian J. Cancer
1994, 31, 192–197. [PubMed]
151. Duffield-Lillico, A.J.; Reid, M.E.; Turnbull, B.W.; Combs, G.F., Jr.; Slate, E.H.; Fischbach, L.A.; Marshall, J.R.;
Clark, L.C. Baseline characteristics and the effect of selenium supplementation on cancer incidence in a
randomized clinical trial: A summary report of the Nutritional Prevention of Cancer Trial. Cancer Epidemiol.
Biomarkers Prev. 2002, 11, 630–639. [PubMed]
152. Li, H.; Stampfer, M.J.; Giovannucci, E.L.; Morris, J.S.; Willett, W.C.; Gaziano, J.M.; Ma, J. A prospective
study of plasma selenium levels and prostate cancer risk. J. Natl. Cancer Inst. 2004, 96, 696–703. [CrossRef]
[PubMed]
153. Overvad, K.; Wang, D.Y.; Olsen, J.; Allen, D.S.; Thorling, E.B.; Bulbrook, R.D.; Hayward, J.L. Selenium in
human mammary carcinogenesis: A case-cohort study. Eur. J. Cancer 1991, 27, 900–902. [CrossRef]
154. Mannisto, S.; Alfthan, G.; Virtanen, M.; Kataja, V.; Uusitupa, M.; Pietinen, P. Toenail selenium and breast
cancer-a case-control study in Finland. Eur. J. Clin. Nutr. 2000, 54, 98–103. [CrossRef] [PubMed]
155. Hardell, L.; Danell, M.; Angqvist, C.A.; Marklund, S.L.; Fredriksson, M.; Zakari, A.L.; Kjellgren, A. Levels of
selenium in plasma and glutathione peroxidase in erythrocytes and the risk of breast cancer. A case-control
study. Biol. Trace Elem. Res. 1993, 36, 99–108. [CrossRef] [PubMed]
156. Hunter, D.J.; Morris, J.S.; Stampfer, M.J.; Colditz, G.A.; Speizer, F.E.; Willett, W.C. A prospective study of
selenium status and breast cancer risk. JAMA 1990, 264, 1128–1131. [CrossRef] [PubMed]
157. Kok, D.E.; Kiemeney, L.A.; Verhaegh, G.W.; Schalken, J.A.; van Lin, E.N.; Sedelaar, J.P.; Witjes, J.A.;
Hulsbergen-van de Kaa, C.A.; van ‘t Veer, P.; Kampman, E.; et al. A short-term intervention with selenium
affects expression of genes implicated in the epithelial-to-mesenchymal transition in the prostate. Oncotarget
2017, 8, 10565–10579. [CrossRef] [PubMed]
158. Radhakrishnan, N.; Dinand, V.; Rao, S.; Gupta, P.; Toteja, G.S.; Kalra, M.; Yadav, S.P.; Sachdeva, A. Antioxidant
levels at diagnosis in childhood acute lymphoblastic leukemia. Indian J. Pediatr. 2013, 80, 292–296. [CrossRef]
[PubMed]
159. Masri, D.S. Microquantity for macroquality: Case study on the effect of selenium on chronic neutropenia.
J. Pediatr. Hematol. Oncol. 2011, 33, e361–e362. [CrossRef] [PubMed]
160. Rocha, K.C.; Vieira, M.L.; Beltrame, R.L.; Cartum, J.; Alves, S.I.; Azzalis, L.A.; Junqueira, V.B.; Pereira, E.C.;
Fonseca, F.L. Impact of Selenium Supplementation in Neutropenia and Immunoglobulin Production in
Childhood Cancer Patients. J. Med. Food 2016, 19, 560–568. [CrossRef] [PubMed]
161. Rohr-Udilova, N.; Bauer, E.; Timelthaler, G.; Eferl, R.; Stolze, K.; Pinter, M.; Seif, M.; Hayden, H.; Reiberger, T.;
Schulte-Hermann, R.; et al. Impact of glutathione peroxidase 4 on cell proliferation, angiogenesis and
cytokine production in hepatocellular carcinoma. Oncotarget 2018, 9, 10054–10068. [CrossRef] [PubMed]
162. Rohr-Udilova, N.; Sieghart, W.; Eferl, R.; Stoiber, D.; Bjorkhem-Bergman, L.; Eriksson, L.C.; Stolze, K.;
Hayden, H.; Keppler, B.; Sagmeister, S.; et al. Antagonistic effects of selenium and lipid peroxides on growth
control in early hepatocellular carcinoma. Hepatology 2012, 55, 1112–1121. [CrossRef] [PubMed]
163. Ren, Y.; Poon, R.T.; Tsui, H.T.; Chen, W.H.; Li, Z.; Lau, C.; Yu, W.C.; Fan, S.T. Interleukin-8 serum levels
in patients with hepatocellular carcinoma: Correlations with clinicopathological features and prognosis.
Clin. Cancer Res. 2003, 9, 5996–6001. [PubMed]
164. Gautam, P.K.; Kumar, S.; Tomar, M.S.; Singh, R.K.; Acharya, A.; Kumar, S.; Ram, B. Selenium nanoparticles
induce suppressed function of tumor associated macrophages and inhibit Dalton’s lymphoma proliferation.
Biochem. Biophys. Rep. 2017, 12, 172–184. [CrossRef] [PubMed]
165. Diwakar, B.T.; Korwar, A.M.; Paulson, R.F.; Prabhu, K.S. The Regulation of Pathways of Inflammation and
Resolution in Immune Cells and Cancer Stem Cells by Selenium. Adv. Cancer Res. 2017, 136, 153–172.
[PubMed]
227
Nutrients 2018, 10, 1203
166. Zhang, J.; Basher, F.; Wu, J.D. NKG2D Ligands in Tumor Immunity: Two Sides of a Coin. Front. Immunol
2015, 6, 97. [CrossRef] [PubMed]
167. Hagemann-Jensen, M.; Uhlenbrock, F.; Kehlet, S.; Andresen, L.; Gabel-Jensen, C.; Ellgaard, L.;
Gammelgaard, B.; Skov, S. The selenium metabolite methylselenol regulates the expression of ligands that
trigger immune activation through the lymphocyte receptor NKG2D. J. Biol. Chem. 2014, 289, 31576–31590.
[CrossRef] [PubMed]
168. Lennicke, C.; Rahn, J.; Bukur, J.; Hochgrafe, F.; Wessjohann, L.A.; Lichtenfels, R.; Seliger, B. Modulation of
MHC class I surface expression in B16F10 melanoma cells by methylseleninic acid. Oncoimmunology 2017, 6,
e1259049. [CrossRef] [PubMed]
169. Rose, A.H.; Bertino, P.; Hoffmann, F.W.; Gaudino, G.; Carbone, M.; Hoffmann, P.R. Increasing dietary
selenium elevates reducing capacity and ERK activation associated with accelerated progression of select
mesothelioma tumors. Am. J. Pathol. 2014, 184, 1041–1049. [CrossRef] [PubMed]
170. Faghfuri, E.; Yazdi, M.H.; Mahdavi, M.; Sepehrizadeh, Z.; Faramarzi, M.A.; Mavandadnejad, F.;
Shahverdi, A.R. Dose-response relationship study of selenium nanoparticles as an immunostimulatory
agent in cancer-bearing mice. Arch. Med. Res. 2015, 46, 31–37. [CrossRef] [PubMed]
171. Wang, H.; Chan, Y.L.; Li, T.L.; Bauer, B.A.; Hsia, S.; Wang, C.H.; Huang, J.S.; Wang, H.M.; Yeh, K.Y.;
Huang, T.H.; et al. Reduction of splenic immunosuppressive cells and enhancement of anti-tumor immunity
by synergy of fish oil and selenium yeast. PLoS ONE 2013, 8, e52912. [CrossRef] [PubMed]
172. Yazdi, M.H.; Mahdavi, M.; Varastehmoradi, B.; Faramarzi, M.A.; Shahverdi, A.R. The immunostimulatory
effect of biogenic selenium nanoparticles on the 4T1 breast cancer model: An in vivo study. Biol. Trace
Elem. Res. 2012, 149, 22–28. [CrossRef] [PubMed]
173. Kim, H.Y. The methionine sulfoxide reduction system: Selenium utilization and methionine sulfoxide
reductase enzymes and their functions. Antioxid. Redox Signal. 2013, 19, 958–969. [CrossRef] [PubMed]
174. Lee, B.C.; Peterfi, Z.; Hoffmann, F.W.; Moore, R.E.; Kaya, A.; Avanesov, A.; Tarrago, L.; Zhou, Y.;
Weerapana, E.; Fomenko, D.E.; et al. MsrB1 and MICALs regulate actin assembly and macrophage function
via reversible stereoselective methionine oxidation. Mol. Cell. 2013, 51, 397–404. [CrossRef] [PubMed]
175. Saha, S.S.; Hashino, M.; Suzuki, J.; Uda, A.; Watanabe, K.; Shimizu, T.; Watarai, M. Contribution of methionine
sulfoxide reductase B (MsrB) to Francisella tularensis infection in mice. FEMS Microbiol. Lett. 2017, 364.
[CrossRef] [PubMed]
176. Fredericks, G.J.; Hoffmann, F.W.; Rose, A.H.; Osterheld, H.J.; Hess, F.M.; Mercier, F.; Hoffmann, P.R.
Stable expression and function of the inositol 1,4,5-triphosphate receptor requires palmitoylation by a
DHHC6/selenoprotein K complex. Proc. Natl. Acad. Sci. USA 2014, 111, 16478–16483. [CrossRef] [PubMed]
177. Fredericks, G.J.; Hoffmann, F.W.; Hondal, R.J.; Rozovsky, S.; Urschitz, J.; Hoffmann, P.R. Selenoprotein
K Increases Efficiency of DHHC6 Catalyzed Protein Palmitoylation by Stabilizing the Acyl-DHHC6
Intermediate. Antioxidants (Basel) 2017, 7, 4. [CrossRef] [PubMed]
178. Meiler, S.; Baumer, Y.; Huang, Z.; Hoffmann, F.W.; Fredericks, G.J.; Rose, A.H.; Norton, R.L.; Hoffmann, P.R.;
Boisvert, W.A. Selenoprotein K is required for palmitoylation of CD36 in macrophages: Implications in foam
cell formation and atherogenesis. J. Leukoc. Biol. 2013, 93, 771–780. [CrossRef] [PubMed]
179. Norton, R.L.; Fredericks, G.J.; Huang, Z.; Fay, J.D.; Hoffmann, F.W.; Hoffmann, P.R. Selenoprotein K
regulation of palmitoylation and calpain cleavage of ASAP2 is required for efficient FcgammaR-mediated
phagocytosis. J. Leukoc. Biol. 2017, 101, 439–448. [CrossRef] [PubMed]
© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
228
nutrients
Review
Human Milk Oligosaccharides and Immune
System Development
Julio Plaza-Díaz 1,2,3 , Luis Fontana 1,2,3 and Angel Gil 1,2,3,4, *
1 Department of Biochemistry and Molecular Biology II, School of Pharmacy, University of Granada,
18071 Granada, Spain; jrplaza@ugr.es (J.P.-D.); fontana@ugr.es (L.F.)
2 Institute of Nutrition and Food Technology “José Mataix”, Biomedical Research Center,
Parque Tecnológico Ciencias de la Salud, University of Granada, Armilla, 18100 Granada, Spain
3 Instituto de Investigación Biosanitaria ibs., 18014 Granada, Spain
4 CIBEROBN, Instituto de Salud Carlos III, 28029 Madrid, Spain
* Correspondence: agil@ugr.es; Tel.: +34-9-5824-6139 or +34-9-5824-1000 (ext. 20307)
Abstract: Maternal milk contains compounds that may affect newborn immunity. Among these
are a group of oligosaccharides that are synthesized in the mammary gland from lactose;
these oligosaccharides have been termed human milk oligosaccharides (HMOs). The amount of
HMOs present in human milk is greater than the amount of protein. In fact, HMOs are the third-most
abundant solid component in maternal milk after lactose and lipids, and are thus considered to
be key components. The importance of HMOs may be explained by their inhibitory effects on the
adhesion of microorganisms to the intestinal mucosa, the growth of pathogens through the production
of bacteriocins and organic acids, and the expression of genes that are involved in inflammation.
This review begins with short descriptions of the basic structures of HMOs and the gut immune
system, continues with the beneficial effects of HMOs shown in cell and animal studies, and it ends
with the observational and randomized controlled trials carried out in humans to date, with particular
emphasis on their effect on immune system development. HMOs seem to protect breastfed infants
against microbial infections. The protective effect has been found to be exerted through cell signaling
and cell-to-cell recognition events, enrichment of the protective gut microbiota, the modulation of
microbial adhesion, and the invasion of the infant intestinal mucosa. In addition, infants fed formula
supplemented with selected HMOs exhibit a pattern of inflammatory cytokines closer to that of
exclusively breastfed infants. Unfortunately, the positive effects found in preclinical studies have
not been substantiated in the few randomized, double-blinded, multicenter, controlled trials that are
available, perhaps partly because these studies focus on aspects other than the immune response
(e.g., growth, tolerance, and stool microbiota).
1. Introduction
Breastfeeding has many beneficial effects in newborns. The relative risks of diarrhea incidence,
diarrhea mortality, pneumonia incidence, and pneumonia mortality are kept to a minimum in
exclusively breastfed infants. These protective effects, although less robust, are also observed in
partially breastfed infants when compared with milk formula-fed infants [1]. However, in addition
to protecting against infection, human milk has both short-term and long-term effects, such as
prevention and protection against allergic reactions; optimal behavioral, cognitive, and gastrointestinal
development; and, may protect against chronic diseases, such as diabetes, obesity, hypertension,
and autoimmune and cardiovascular diseases [2]. The long-term effects of human milk are related to
so-called early programming [3].
Human milk contains many bioactive compounds that may affect immunity (e.g., cytokines,
growth factors, hormones, digestive enzymes, transporters, and antimicrobial factors). The latter
category of antimicrobial factors includes glycans, among which exists a group of oligosaccharides with
different structures that are synthesized from lactose in the mammary gland. These oligosaccharides
have been termed human milk oligosaccharides (HMOs). In addition, human milk contains probiotics,
which reside in the microbiota of the breast tissue and may also have a role in neonate immunity [4,5].
Although HMOs were originally described and referred to as “gynolactose” by Lespagnol and
Polonowski in 1930 [4], they have attracted considerable attention in recent years because of their
biological roles. The HMO fraction (5–15 g/L) that is present in human milk is greater than the protein
fraction (8 g/L). In fact, HMOs are the third-most abundant solid component in maternal milk after
lactose (70 g/L) and lipids (40 g/L), and accordingly, they are considered to be key compounds [4,5].
HMOs have been described to inhibit (i) the adhesion of microorganisms to the intestinal mucosa;
(ii) the growth of pathogens through the production of bacteriocins and organic acids; and (iii) the
expression of genes involved in inflammation [4–6]. Although many studies regarding the composition
of oligosaccharides in human milk have been published, there are few publications about the roles of
these compounds in general and in immunity in particular [5,7]. The aim of this review article was to
fill this gap by surveying the in vitro and in vivo effects of HMOs, focusing mainly on immunity.
(a) Neutral (fucosylated) HMOs are neutral and contain fucose at the terminal position
(e.g., 2 -fucosyllactose (2 -FL) and lactodifucopentaose). They represent 35% to 50% of the
total HMO content.
(b) Neutral N-containing (nonfucosylated) HMOs are neutral, contain N-acetylglucosamine at the
terminal position (e.g., lacto-N-tetraose), and represent 42% to 55% of the total HMO content.
Neutral HMOs account for more than 75% of the total HMOs in human breast milk.
(c) Acid (sialylated) HMOs are acidic and contain sialic acid at the terminal position
(e.g., 2 -sialyllactose). They represent 12% to 14% of the total HMO content.
230
Nutrients 2018, 10, 1038
The amount and composition of HMOs vary among women. The HMO composition is determined
genetically and mirrors blood group characteristics, which depend on the expression of certain
glycosyltransferases. Four milk groups can be assigned based on the Secretor (Se) and Lewis (Le) blood
group system, which is determined by the activity of two gene loci encoding α1-2-fucosyltransferase
(FUT2, encoded by the Se gene) and α1-3/4-fucosyltransferase (FUT3, encoded by the Le gene) [6–17].
Individuals with an active Se locus are classified as “secretors”. The milk of secretor women is abundant
in 2 -FL, lacto-N-fucopentaose I (LNFP I), and other α1-2-fucosylated HMOs. In contrast, non-secretor
women lack a functional FUT2 enzyme, and their milk does not contain α1-2-fucosylated HMOs.
Individuals with an active Le locus are classified as Le-positive. They express FUT3, which transfers
Fuc with a α1-4 linkage to subterminal GlcNAc on type 1 chains. In contrast, the milk of Le-negative
women lacks these specific α1-4-fucosylated HMOs, e.g., LNFP II [9,10,14]. Therefore, breast milk can
be assigned to one of the four groups based on the expression of FUT2 and FUT3: Le-positive secretors
(Le+Se+), Le-negative secretors (Le-Se+), Le-positive nonsecretors (Le+Se-), and Le-negative nonsecretors
(Le-Se+) (Table 1) [9,10,14]:
This classification, however, is an oversimplification. FUT2 and FUT3 compete for some of the
same substrates [19–21], and the levels of enzyme expression and activity translate into different
profiles throughout the population. Even the milk of Le-negative nonsecretor women who express
neither FUT2 nor FUT3 contains fucosylated HMOs, such as 3FL or LNFP III, suggesting that other Se-
231
Nutrients 2018, 10, 1038
and Le-independent FUTs may be involved [14,22]. In addition, α1-2-fucosylated HMOs have been
found in the milk of nonsecretor women near the end of lactation, and Newburg et al. suggested that
FUT1 might also participate in HMO fucosylation [22]. In addition, fucosylation in preterm milk is
not as well regulated as in term milk, resulting in higher within and between mother variation in
women delivering preterm vs term. In fact, of particular clinical interest, the α1,2-linked fucosylated
oligosaccharide 2 -fucosyllactose, which is an indicator of secretor status, is not consistently present
across the lactation of several mothers that delivered preterm [23].
The amount and composition of HMOs also vary over the course of lactation. Whereas, colostrum
contains as much as 20–25 g/L of HMOs, as milk production matures, HMO concentrations decline to
5–20 g/L, which still exceeds the concentration of total milk protein [4]. The milk of mothers delivering
preterm infants has higher HMO concentrations than term milk [14], whereas preterm milk contains
lower levels of fucosylated HMOs than term milk [24], and no differences in neutral and acidic HMOs
are found between preterm and term milk [25].
Figure 2. Main lymphocyte populations of the gut-associated lymphoid tissue (GALT). Modified with
permission from [26].
232
Nutrients 2018, 10, 1038
The antigens that are present in the intestinal lumen are processed and transported into the
Peyer patches via the M cells, which are located among the enterocytes in the epithelium. Once in the
Peyer patches, the antigens interact with antigen-presenting cells (APCs), which are responsible for
presenting those antigens to immature B and T lymphocytes that are residing in both the germinal
centers and interfollicular regions. After their activation by antigens, the immature B and T cells
drain down the lymph nodes and migrate through the thoracic duct to the bloodstream. They may
circulate for a few days and later differentiate into mature effector cells that migrate to the lamina
propria or memory cells, which again travel to the Peyer patches [26,27]. The so-called dendritic cells
that are present in the Peyer patches and the lamina propria have been shown to form pseudopods
and interact directly with antigens that are present in the intestinal lumen, after which they process
the antigens and present them to other underlying cell lineages without the involvement of the
M cells [28,29]. Another population of effector cells consisting of IELs may interact with antigens
entering the gastrointestinal tract without following the course mentioned above. In recent years,
a new type of cells, innate lymphoid cells (ILCs), have been described along with their functions [30].
ILCs are present in the intestine and other mucosae and participate in tissue homeostasis, inflammation,
and autoimmunity, although their main function is the development of the gut barrier.
The potential beneficial effects of HMOs might be related to their capacity to interact with a
number of receptors that are located in intestinal immune cells [31].
233
Nutrients 2018, 10, 1038
Fucosylated HMOs have been reported to inhibit (i) the binding of several pathogens, such as
Campylobacter jejuni [46], Norwald-like virus [47], and Helicobacter pylori [48], and (ii) the heat-stable
enterotoxin of Escherichia coli [49] to intestinal cells.
The addition of HMOs was tested in T84 cell membranes to establish the inhibition of
enterotoxin-producing Escherichia coli. The administration of HMOs repressed E. coli guanylate cyclase
activity and cyclic GMP production in these cells [50]. Uropathogenic E. coli strains expressing P (Pap)
and P-like (Prs) fimbriae are responsible for infections of the urinary tract. The hemagglutination
that is mediated by these strains was inhibited by HMOs, especially by the sialylated fraction [51].
Fractions of HMOs were evaluated for their ability to inhibit the adhesion of E. coli serotype O119,
Vibrio cholerae, and Salmonella fyris in differentiated Caco-2 cells. The evaluated HMOs inhibited the
adhesion of these pathogens to epithelial cells [52]. Oligosaccharides from milk might block the action
of PA-IIL, a fucose-binding lectin of the human pathogen Pseudomonas aeruginosa, through competition
for the receptor and further binding [53]. In particular, a significant reduction in uropathogenic E. coli
internalization into HMO-pretreated epithelial cells was detected without observing any binding
to these cells [54]. HMOs from pooled human milk significantly reduced enteropathogenic E. coli
strain 2348/69 (serotype O127:H6) attachment to cultured epithelial cells [55]. Likewise, treatment
with HMOs reduced the invasion of human premature intestinal epithelial cells by C. albicans in a
dose-dependent manner [56].
Colonization and invasion require the attachment of trophozoites to the host’s mucosa.
HMOs reduce E. histolytica attachment and cytotoxicity; in fact, pooled HMOs detach E. histolytica
by more than 80%; moreover, HMOs rescue E. histolytica-induced destruction of human intestinal
epithelial HT-29 cells in a dose-dependent manner [57].
234
Nutrients 2018, 10, 1038
HMOs also directly modulate immune responses. HMOs may act either locally, on cells of the
mucosa-associated lymphoid tissues, or at a systemic level, as 1% of HMOs are absorbed and reach the
systemic circulation [4,72–74].
The transcriptional response of colonic epithelial cells that are treated with HMOs was investigated
in HT-29 cells. The expression of several cytokines (such as IL-1β, IL-8, colony-stimulating factor 2,
IL-17C and platelet factor 4 (PF4)), chemokines (such as CXCL1, CXCL3, CXCL2, CXCL6, CCL5, CCL20,
and CX3CL1), and cell surface receptors (interferon γ receptor 1, IFNGR1), intercellular adhesion
molecule-1 (ICAM-1), intercellular adhesion molecule-2 (ICAM-2), and IL-10 receptor a (IL10RA)
in HT-29 cells was influenced by the administration of HMOs [75]. The aforementioned cytokines,
chemokines, and cell surface receptors are implicated in the development and maturation of the
intestinal immune response [75].
Using a cellular model with intestinal epithelial cells (T84/HCT8/FHs74) and HeLa cells,
He et al. [76] investigated the effects of HMOs from colostrum in fetal human intestinal mucosa
cells. These authors identified networks controlling immune cell communication, intestinal mucosal
immune system differentiation, and homeostasis. HMOs treatment decreased cytokine protein levels,
such as IL-8, IL-6, monocyte chemoattractant protein-1/2 and IL-1β, while increasing the levels of
cytokines that are involved in tissue repair and homeostasis [76].
Gram-negative pathogenic bacteria might activate mucosal inflammation through the binding of
LPS to intestinal toll-like receptor 4 (TLR4) in epithelial intestinal cells (IECs). Under in vitro conditions,
IECs were treated with a strain of enterotoxigenic E. coli to evaluate the inhibitory effect of HMO
treatment on the secretion of IL-8. Both treatments (a mixture of HMOs and 2 -FL alone) successfully
decreased the LPS-dependent stimulation of IL-8 through the attenuation of CD14 induction.
CD14 expression mediates the LPS-TLR4 stimulation of portions of the “macrophage migration
inhibitory factors” inflammatory pathway via suppressors of cytokine signaling 2/signal transducer
and the activator of transcription (STAT) 3/NF-κB [76]. The effects of different oligosaccharide fractions
on leukocyte rolling and adhesion were determined. Two active compounds (3 -sialyl-lactose and
3 -sialyl-3-fucosyl-lactose) exhibit an inhibitory effect on leukocyte rolling and adhesion, decreasing
the incidence of inflammatory diseases due to their anti-inflammatory activities [77].
The administration of HMOs also demonstrated growth inhibition of Streptococcus agalactiae
(group B Streptococcus). This bacteriostatic activity was mediated through the action of a putative
glycosyltransferase that confers resistance to oligosaccharides [78]. The human epithelial cell lines
HEp-2 and HT-29 were infected with C. jejuni 81-176 and were treated with 2 FL to evaluate the degree
of infection and inflammatory response. Treatment with 2 -FL attenuated the majority of C. jejuni
invasion and decreased the release of IL-8 and IL-1b by 80–90% as well as decreasing the level of the
neutrophil chemoattractant macrophage inflammatory protein 2 (MIP-2) [79].
Bacterial strains are not the only pathogens that are inhibited by the action of HMOs. In addition,
some evidence suggests that HMOs act against viruses [6]. In particular, 2 -FL and 3-FL can
structurally mimic histo-blood group antigens and block the binding of norovirus, which can
cause acute gastroenteritis in humans [80]. The effects of 2 FL, 6 -sialyllactose, 3 -sialyllactose and
lacto-N-neotetraose on peripheral blood mononuclear cells (PBMCs) following respiratory viral
infection were investigated in vitro. The administration of 2 -FL significantly decreased the respiratory
syncytial viral load and cytokines that are associated with disease severity (IL-6, IL-8, MIP-1α) and
inflammation (TNF-α, MCP-1) in airway epithelial cells. Lacto-N-neotetraose and 6 -sialyllactose
treatments decreased the influenza viral load in airway epithelial cells, and only 6 -sialyllactose
decreased CXCL10 and TNF-α in respiratory syncytial virus-infected PBMCs [81]. Particularly,
HMOs containing more than one unit of fucose may exhibit stronger binding capacities when compared
with single fucose HMOs [82].
235
Nutrients 2018, 10, 1038
236
Nutrients 2018, 10, 1038
cow’s milk allergy, they are not required to prevent cow’s milk allergy. Therefore, other mechanisms
must be in play [90].
HMOs have multiple immunomodulatory functions that influence child´s health [91]. Kuhn et al.
evaluated the effects of HMO from breast milk on the survival of uninfected children born to
HIV-infected mothers. Higher maternal breast milk concentrations of 2-linked fucosylated HMOs
(2 -FL and lacto-N-fucopentaose I, as well as 3FL and lacto-N-fucopentaose II/III) were significantly
associated with reduced mortality [91].
Breast milk samples were analyzed to determine the levels of 2-linked fucosyloligosaccharides
and their relationship with the incidence of moderate-to-severe diarrhea. The incidence of diarrhea was
lower in infants fed milk containing high levels of total 2-linked fucosyloligosaccharides. Campylobacter,
and calicivirus-induced diarrhea occurred less often in infants whose mother’s milk contained high
levels of 2 -FL and lacto-N-difucohexaose [92].
237
Nutrients 2018, 10, 1038
multicenter, double-blinded, controlled study involving 131 full-term infants enrolled between 0 and
8 days of age [95]. Infants were randomly allocated to receive either milk-based infant formula not
containing oligosaccharides (n = 42) or milk-based infant formula containing 2 -FL (0.2 g/L) and
scFOS (2 g/L) (n = 46). A group of 43 breastfed infants were also included. The intervention was
performed for 35 days. The primary outcome was the average mean rank stool consistency (MRSC)
from study day 1 to visit 3, calculated from stool records. From study day 1 to visit 3, no difference in
stool consistency was observed. At visit 3, there were no differences between groups in the average
volume of study formula intake, the number of study formula feedings per day, anthropometric data,
or percentage of feedings with spitting up or vomiting. The conclusion of this study was that the
formula containing 2 FL was well tolerated in infants as evidenced by stool consistency, formula intake,
anthropometric data and percent feedings with spitup/vomit similar to that of infants fed a formula
without oligosaccharides or breast milk [95].
The effects of feeding infant formulas supplemented with two human milk-identical oligosaccharides,
2 FL, and lacto-N-neotetraose (LNnT), on infant growth, tolerability, gut microbiota, and medication
use were investigated in a randomized, controlled, multicenter trial in Italy and Belgium [96–98].
One hundred and seventy-five healthy, full-term infants were randomly allocated after birth (between
0–14 days of age) to one of the two formula feeding groups: intact cow’s milk-based whey-predominant
infant formula with the addition of 2 -FL (1.0 g/L) and LNnT (0.5 g/L) (test group, n = 88) or intact
cow’s milk-based whey-predominant infant formula (control group, n = 87). The formulas were given
up to six months of age (exclusive formula feeding up to four months). A group of 38 exclusively
breastfed infants were enrolled at three months as a reference (breastfed group). The primary endpoint
was weight gain through four months of age. The secondary endpoints were anthropometry, stool
characteristics, stool microbiota (at 3 and 12 months of age, obtained using 16S rRNA gene sequencing
and metagenomics), stool metabolic signature (at 3 and 12 months of age, obtained using proton
NMR-based metabolic profiling,), digestive tolerance, and morbidity (reported by parents) through
12 months of age. The weight gain up to four months of age of infants fed formula supplemented with
2 -FL and LNnT was not inferior to the weight gain of infants fed unsupplemented formula. The mean
weight, length, head circumference, and BMI up to 12 months of age of infants fed formulas with or
without 2 -FL and LNnT were close to the WHO Growth Standards and did not differ between the two
groups. Digestive tolerance was similar between the two groups. Infants receiving formula containing
2 -FL and LNnT had significantly fewer parental reports of lower respiratory tract infections (19.3% vs.
34.5%; OR 0.45, 95% CI 0.21–0.95; p = 0.027), particularly bronchitis (10.2% vs. 27.6%; OR 0.30, 95% CI
0.11–0.73; p = 0.004), up to 12 months of age (42% vs. 60.9%; OR 0.47, 95% CI 0.24–0.89; p = 0.016),
and lower medication (antibiotics up to 12 months of age, antipyretics (15.9% vs. 29.9%; OR 0.44; 95%
CI 0.2–0.98; p = 0.032) up to four months of age) than infants fed formula without 2 FL and LNnT [96].
In conclusion, feeding infant formula containing two HMOs (2 FL and LNnT) during the first six
months of age are safe, well-tolerated, and supports age-appropriate growth [81]. Also, the observed
effects on reduced morbidity and medication use in infants up to 12 months of age, when feeding
formula with HMOs, suggest that 2 FL and LNnT may provide immune benefits [81].
In a second work, these authors reported the effects of 2 -FL and LNnT on the infant gut
microbiota [97]. The microbiota composition of infants that were fed formula containing 2 -FL and
LNnT was significantly different from that of infants fed nonsupplemented formula (p < 0.001) at
the genus level and closer to that of breastfed infants at three months of age. Three main bacterial
genera (Bifidobacterium, Escherichia, and Peptostreptococcaceae) showed significant differences in infants
fed formula with or without 2 ‘FL and LNnT at three months of age: greater abundance of beneficial
Bifidobacterium (p < 0.01); and, lower abundance of potentially pathogenic Escherichia (p < 0.01) as well
as of unclassified Peptostreptococcaceae (p < 0.05) were observed in infants fed formula containing 2 -FL
and LNnT as compared to infants fed nonsupplemented formula. These values were closer to the levels
that were observed in breastfed infants. The biochemical composition of the stools was explored by the
quantitative profiling of major metabolites to gain additional information on the compositional aspects.
238
Nutrients 2018, 10, 1038
The stool contents of some amino acids (phenylalanine, tyrosine, isoleucine), some SCFAs (propionate,
butyrate), and some organic acids (lactate) in infants fed formula with 2 -FL and LNnT tended to be
closer to those that were observed in breastfed infants than those in infants fed nonsupplemented
formula [97]. This study shows that a formula supplemented with 2 FL and LNnT shifts stool microbiota
and metabolic signatures of infants born at term closer to that of breastfed infants [97].
Finally, in a third work, the same authors found that, at three months, the microbiota composition
in the test group appeared closer to that of the breastfed group than to that of the control group
according to alpha (within group) and beta (between groups) diversity analyses of the microbiota
and the distribution of microbiota community types (A, B, and C). Supplementation with both HMOs
decreased the number of infants with formula-specific C-community (fecal community type/FCT C)
and increased those with the breastfed-specific B-community (FCT B). Cumulative antibiotic use up to
12 months was associated with the FCT distribution at three months. Infants with FCT B at 3 months
were less likely to be treated with antibiotics (OR 0.4 (95% CI, 0.17–0.93; p = 0.033)), while infants
with FCT C were more likely to be treated with antibiotics during the first 12 months (OR 3.3 (95% CI,
1.54–7.02; p = 0.0025)). The microbiota community type at three months was not associated with other
parent-reported infection-related morbidities [98]. This study confirms the microbiota results of the
previous one [97] and shows that infants with a breastfed specific microbiota community type (FCT B)
are less likely to need antibiotics.
239
Nutrients 2018, 10, 1038
HMO-supplemented formulas but not to generate actual evidence of the potential preventive effects of
HMOs against infectious diseases. Therefore, new RCT studies in infants with the appropriate power
should be designed to ascertain the roles of HMOs in the prevention of diarrhea, pneumonia, and other
respiratory diseases.
Author Contributions: All three authors participated in the bibliographic search, discussion and writing of
the manuscript.
Funding: This research received no external funding.
Acknowledgments: Julio Plaza-Diaz and Angel Gil are part of University of Granada, Plan Propio de Investigación
2016, Excellence actions: Units of Excellence; Unit of Excellence on Exercise and Health (UCEES).
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Black, R.E.; Allen, L.H.; Bhutta, Z.A.; Caulfield, L.E.; de Onis, M.; Ezzati, M.; Mathers, C.; Rivera, J.; Maternal
and Child Undernutrition Study Group. Maternal and child undernutrition: Global and regional exposures
and health consequences. Lancet 2008, 371, 243–260. [CrossRef]
2. Dieterich, C.M.; Felice, J.P.; O’Sullivan, E.; Rasmussen, K.M. Breastfeeding and health outcomes for the
mother-infant dyad. Pediatr. Clin. N. Am. 2013, 60, 31–48. [CrossRef] [PubMed]
3. Koletzko, B.; Brands, B.; Grote, V.; Kirchberg, F.F.; Prell, C.; Rzehak, P.; Uhl, O.; Weber, M. Early Nutrition
Programming Project. Long-term health impact of early nutrition: The power of programming. Ann. Nutr. Metab.
2017, 70, 161–169. [CrossRef] [PubMed]
4. Bode, L. Human milk oligosaccharides: Every baby needs a sugar mama. Glycobiology 2012, 22, 1147–1162.
[CrossRef] [PubMed]
5. Musilova, S.; Rada, V.; Vlkova, E.; Bunesova, V. Beneficial effects of human milk oligosaccharides on gut
microbiota. Benef. Microbes 2014, 5, 273–283. [CrossRef] [PubMed]
6. Morozov, V.; Hansman, G.; Hanisch, F.G.; Schroten, H.; Kunz, C. Human milk oligosaccharides as promising
antivirals. Mol. Nutr. Food Res. 2018, 62, e1700679. [CrossRef] [PubMed]
7. Doherty, A.M.; Lodge, C.J.; Dharmage, S.C.; Dai, X.; Bode, L.; Lowe, A.J. Human milk oligosaccharides and
associations with immune-mediated disease and infection in childhood: A Systematic Review. Front. Pediatr.
2018, 6, 91. [CrossRef] [PubMed]
8. Zivkovic, A.M.; German, J.B.; Lebrilla, C.B.; Mills, D.A. Human milk glycobiome and its impact on the infant
gastrointestinal microbiota. Proc. Natl. Acad. Sci. USA 2011, 108, 4653–4658. [CrossRef] [PubMed]
9. Bode, L. The functional biology of human milk oligosaccharides. Early Hum. Dev. 2015, 91, 619–622.
[CrossRef] [PubMed]
10. Smilowitz, J.; Lebrilla, C.; Mills, D.; German, J.; Freeman, S. Breast milk oligosaccharides: Structure-function
relationships in the neonate. Annu. Rev. Nutr. 2014, 34, 143–169. [CrossRef] [PubMed]
11. Kunz, C.; Rudloff, S.; Baier, W.; Klein, N.; Strobel, S. Oligosaccharides in human milk: Structural, functional,
and metabolic aspects. Annu. Rev. Nutr. 2000, 20, 699–722. [CrossRef] [PubMed]
12. Kobata, A. Structures and application of oligosaccharides in human milk. Proc. Jpn. Acad. Ser. B Phys. Biol. Sci.
2010, 86, 731–747. [CrossRef] [PubMed]
13. Jantscher-Krenn, E.; Bode, L. Human milk oligosaccharides and their potential benefits for the breast-fed
neonate. Minerva Pediatr. 2012, 64, 83–99. [PubMed]
14. Bode, L.; Jantscher-Krenn, E. Structure-function relationships of human milk oligosaccharides. Adv. Nutr.
2012, 3, 383S–391S. [CrossRef] [PubMed]
15. Goehring, K.C.; Kennedy, A.D.; Prieto, P.A.; Buck, R.H. Direct evidence for the presence of human milk
oligosaccharides in the circulation of breastfed infants. PLoS ONE 2014, 9, e101692. [CrossRef] [PubMed]
16. Blank, D.; Dotz, V.; Geyer, R.; Kunz, C. Human milk oligosaccharides and Lewis blood group: Individual
high-throughput sample profiling to enhance conclusions from functional studies. Adv. Nutr. 2012, 3,
440S–449S. [CrossRef] [PubMed]
17. Austin, S.; De Castro, C.A.; Bénet, T.; Hou, Y.; Sun, H.; Thakkar, S.K.; Vinyes-Pares, G.; Zhang, Y.; Wang, P.
Temporal change of the content of 10 oligosaccharides in the milk of chinese urban mothers. Nutrients 2016,
8, 346. [CrossRef] [PubMed]
240
Nutrients 2018, 10, 1038
18. Kunz, C.; Meyer, C.; Collado, M.C.; Geiger, L.; Garcia-Mantrana, I.; Bertua-Rios, B.; Martinez-Costa, C.;
Borsch, C.; Rudloff, S. Influence of gestational age, secretor, and lewis blood group status on the
oligosaccharide content of human milk. J. Pediatr. Gastroenterol. Nutr. 2017, 64, 789–798. [CrossRef] [PubMed]
19. Kumazaki, T.; Yoshida, A. Biochemical evidence that secretor gene, Se, is a structural gene encoding a specific
fucosyltransferase. Proc. Natl. Acad. Sci. USA 1984, 81, 4193–4197. [CrossRef] [PubMed]
20. Johnson, P.H.; Watkins, W.M. Purification of the Lewis blood-group gene associated alpha-3/4-fucosyltransferase
from human milk: An enzyme transferring fucose primarily to type 1 and lactose-based oligosaccharide
chains. Glycoconj. J. 1992, 9, 241–249. [CrossRef] [PubMed]
21. Xu, Z.; Vo, L.; Macher, B.A. Structure-function analysis of human α1,3-fucosyltransferase. Amino acids
involved in acceptor substrate specificity. J. Biol. Chem. 1996, 271, 8818–8823. [CrossRef] [PubMed]
22. Newburg, D.S.; Ruiz-Palacios, G.M.; Morrow, A.L. Human milk glycans protect infants against enteric
pathogens. Annu. Rev. Nutr. 2005, 25, 37–58. [CrossRef] [PubMed]
23. De Leoz, M.L.; Gaerlan, S.C.; Strum, J.S.; Dimapasoc, L.M.; Mirmiran, M.; Tancredi, D.J.; Smilowitz, J.T.;
Kalanetra, K.M.; Mills, D.A.; German, J.B.; et al. Lacto-N-tetraose, fucosylation, and secretor status are
highly variable in human milk oligosaccharides from women delivering preterm. J. Proteome Res. 2012, 11,
4662–4672. [CrossRef] [PubMed]
24. Davidson, B.; Meinzen-Derr, J.K.; Wagner, C.L.; Newburg, D.S.; Morrow, A.L. Fucosylated oligosaccharides
in human milk in relation to gestational age and stage of lactation. Adv. Exp. Med. Biol. 2004, 554, 427–430.
[CrossRef] [PubMed]
25. Dotz, V.; Adam, R.; Lochnit, G.; Schroten, H.; Kunz, C. Neutral oligosaccharides in feces of breastfed and
formula-fed infants at different ages. Glycobiology 2016, 26, 1308–1316. [CrossRef] [PubMed]
26. Rueda-Cabrera, R.; Gil, A. Nutrición en inmunidad en el estado de salud. In Tratado de Nutrición; Editorial
Médica Panamericana: Madrid, Spain, 2017; Volume 4, ISBN 9788491101932.
27. Rumbo, M.; Schiffrin, E.J. Ontogeny of intestinal epithelium immune functions: Developmental and
environmental regulation. Cell. Mol. Life Sci. 2005, 62, 1288–1296. [CrossRef] [PubMed]
28. Coombes, J.L.; Powrie, F. Dendritic cells in intestinal immune regulation. Nat. Rev. Immunol. 2008, 6, 411–420.
[CrossRef] [PubMed]
29. Gil, A.; Rueda, R. Interaction of early diet and the development of the immune system. Nutr. Res. Rev. 2002,
15, 263–292. [CrossRef] [PubMed]
30. Klose, C.S.; Artis, D. Innate lymphoid cells as regulators of immunity, inflammation and tissue homeostasis.
Nat. Immunol. 2016, 17, 765–774. [CrossRef] [PubMed]
31. Hardy, H.; Harris, J.; Lyon, E.; Beal, J.; Foey, A.D. Probiotics, prebiotics and immunomodulation of gut
mucosal defences: Homeostasis and immunopathology. Nutrients 2013, 5, 1869–1912. [CrossRef] [PubMed]
32. Pannaraj, P.; Li, F.; Cerini, C.; Bender, J.; Yang, S.; Rollie, A.; Adisetiyo, H.; Zabih, S.; Lincez, P.J.; Bittinger, K.;
et al. Association between breast milk bacterial communities and establishment and development of the
infant gut microbiome. JAMA Pediatr. 2017, 171, 647–654. [CrossRef] [PubMed]
33. Fernández, L.; Langa, S.; Martín, V.; Maldonado, A.; Jiménez, E.; Martín, R.; Rodríguez, J.M. The human
milk microbiota: Origin and potential roles in health and disease. Pharmacol. Res. 2013, 69, 1–10. [CrossRef]
[PubMed]
34. Putignani, L.; Del Chierico, F.; Petrucca, A.; Vernocchi, P.; Dallapiccola, B. The human gut microbiota:
A dynamic interplay with the host from birth to senescence settled during childhood. Pediatr. Res. 2014, 76,
2–10. [CrossRef] [PubMed]
35. Hotamisligil, G.S.; Peraldi, P.; Budavari, A.; Ellis, R.; White, M.F.; Spiegelman, B.M. IRS-1-mediated inhibition
of insulin receptor tyrosine kinase activity in TNF-alpha- and obesity-induced insulin resistance. Science
1996, 271, 665–668. [CrossRef] [PubMed]
36. Bouloumié, A.; Curat, C.A.; Sengenès, C.; Lolmède, K.; Miranville, A.; Busse, R. Role of macrophage tissue
infiltration in metabolic diseases. Curr. Opin. Clin. Nutr. Metab. Care 2005, 8, 347–354. [CrossRef] [PubMed]
37. Cani, P.D.; Amar, J.; Iglesias, M.A.; Poggi, M.; Knauf, C.; Bastelica, D.; Neyrinck, A.M.; Fava, F.; Tuohy, K.M.;
Chabo, C.; et al. Metabolic endotoxemia initiates obesity and insulin resistance. Diabetes 2007, 56, 1761–1772.
[CrossRef] [PubMed]
38. Kunz, C.; Rudloff, S. Biological functions of oligosaccharides in human milk. Acta Pediatr. 1993, 82, 903–912.
[CrossRef]
39. Wold, A.E.; Hanson, L.A. Defence factors in human milk. Curr. Opin. Gastroenterol. 1994, 10, 652–658. [CrossRef]
241
Nutrients 2018, 10, 1038
40. Zopf, D.; Roth, S. Oligosaccharide anti-infective agents. Lancet 1996, 347, 1017–1021. [CrossRef]
41. Sela, D.A.; Chapman, J.; Adeuya, A.; Kim, J.; Chen, F.; Whitehead, T.; Lapidus, A.; Rokhsar, D.; Lebrilla, C.;
German, J. The genome sequence of Bifidobacterium longum subsp. infantis reveals adaptations for milk utilization
within the infant microbiome. Proc. Natl. Acad. Sci. USA 2008, 105, 18964. [CrossRef] [PubMed]
42. Barboza, M.; Pinzon, J.; Wickramasinghe, S.; Froehlich, J.W.; Moeller, I.; Smilowitz, J.T.; Ruhaak, L.R.; Huang, J.;
Lonnerdal, B.; German, J.B.; et al. Glycosylation of human milk lactoferrin exhibits dynamic changes during
early lactation enhancing its role in pathogenic bacteria-host interactions. Mol. Cell. Proteomics 2012, 11,
M111.015248. [CrossRef] [PubMed]
43. Chichlowski, M.; De Lartigue, G.; German, J.B.; Raybould, H.E.; Mills, D.A. Bifidobacteria isolated
from infants and cultured on human milk oligosaccharides affect intestinal epithelial function. J. Pediatr.
Gastroenterol. Nutr. 2012, 55, 321–327. [CrossRef] [PubMed]
44. Newburg, D.S. Do the binding properties of oligosaccharides in milk protect human infants from
gastrointestinal bacteria? J. Nutr. 1997, 127, 980S. [CrossRef] [PubMed]
45. Varki, A. Biological roles of oligosaccharides: All of the theories are correct. Glycobiology 1993, 3, 97–130.
[CrossRef] [PubMed]
46. Ruiz-Palacios, G.M.; Cervantes, L.E.; Ramos, P.; Chavez-Munguia, B.; Newburg, D.S. Campylobacter jejuni
binds intestinal H(O) antigen (Fuc alpha 1, 2Gal beta 1, 4GlcNAc), and fucosyloligosaccharides of human
milk inhibit its binding and infection. J. Biol Chem. 2003, 278, 14112–14120. [CrossRef] [PubMed]
47. Huang, P.; Farkas, T.; Marionneau, S.; Zhong, W.; Ruvoen-Clouet, N.; Morrow, A.L.; Altaye, M.;
Pickering, L.K.; Newburg, D.S.; Le Pendu, J.; et al. Noroviruses bind to human ABO, Lewis, and secretor
histo-blood group antigens: Identification of 4 distinct strain-specific patterns. J. Infect. Dis. 2003, 188, 19–31.
[CrossRef] [PubMed]
48. Xu, H.T.; Zhao, Y.F.; Lian, Z.X.; Fan, B.L.; Zhao, Z.H.; Yu, S.Y.; Dai, Y.P.; Wang, L.L.; Niu, H.L.; Li, N.;
et al. Effects of fucosylated milk of goat and mouse on Helicobacter pylori binding to Lewis b antigen.
World J. Gastroenterol. 2004, 10, 2063–2066. [CrossRef] [PubMed]
49. Newburg, D.S.; Pickering, L.K.; McCluer, R.H.; Cleary, T.G. Fucosylated oligosaccharides of human milk
protect suckling mice from heat-stabile enterotoxin of Escherichia coli. J. Infect. Dis. 1990, 162, 1075–1080.
[CrossRef] [PubMed]
50. Crane, J.K.; Azar, S.S.; Stam, A.; Newburg, D.S. Oligosaccharides from human milk block binding and
activity of the Escherichia coli heat-stable enterotoxin (STa) in T84 intestinal cells. J. Nutr. 1994, 124, 2358–2364.
[CrossRef] [PubMed]
51. Martín-Sosa, S.; Martín, M.J.; Hueso, P. The sialylated fraction of milk oligosaccharides is partially responsible
for binding to enterotoxigenic and uropathogenic Escherichia coli human strains. J. Nutr. 2002, 132, 3067–3072.
[CrossRef] [PubMed]
52. Coppa, G.V.; Zampini, L.; Galeazzi, T.; Facinelli, B.; Ferrante, L.; Capretti, R.; Orazio, G. Human milk
oligosaccharides inhibit the adhesion to Caco-2 cells of diarrheal pathogens: Escherichia coli, Vibrio cholerae,
and Salmonella fyris. Pediatr. Res. 2006, 59, 377–382. [CrossRef] [PubMed]
53. Perret, S.; Sabin, C.; Dumon, C.; Pokorná, M.; Gautier, C.; Galanina, O.; Ilia, S.; Bovin, N.; Nicaise, M.;
Desmadril, M.; et al. Human milk oligosaccharides shorten rotavirus-induced diarrhea and modulate piglet
mucosal immunity and colonic microbiota. ISME J. 2014, 8, 1609–1620. [CrossRef]
54. Lin, A.E.; Autran, C.A.; Espanola, S.D.; Bode, L.; Nizet, V. Human milk oligosaccharides protect bladder
epithelial cells against uropathogenic Escherichia coli invasion and cytotoxicity. J. Infect. Dis. 2014, 209,
389–398. [CrossRef] [PubMed]
55. Manthey, C.F.; Autran, C.A.; Eckmann, L.; Bode, L. Human milk oligosaccharides protect against
enteropathogenic Escherichia coli attachment in vitro and EPEC colonization in suckling mice. J. Pediatr.
Gastroenterol. Nutr. 2014, 58, 165–168. [CrossRef] [PubMed]
56. Gonia, S.; Tuepker, M.; Heisel, T.; Autran, C.; Bode, L.; Gale, C.A. Human milk oligosaccharides inhibit
candida albicans invasion of human premature intestinal epithelial cells. J. Nutr. 2015, 145, 1992–1998.
[CrossRef] [PubMed]
57. Jantscher-Krenn, E.; Lauwaet, T.; Bliss, L.A.; Reed, S.L.; Gillin, F.D.; Bode, L. Human milk oligosaccharides
reduce Entamoeba histolytica attachment and cytotoxicity in vitro. Br. J. Nutr. 2012, 108, 1839–1846. [CrossRef]
[PubMed]
242
Nutrients 2018, 10, 1038
58. Canfora, E.E.; Jocken, J.W.; Blaak, E.E. Short-chain fatty acids in control of body weight and insulin sensitivity.
Nat. Rev. Endocrinol. 2015, 11, 577–591. [CrossRef] [PubMed]
59. Hur, K.Y.; Lee, M.S. Gut Microbiota and Metabolic Disorders. Diabetes Metab. J. 2015, 39, 198–203. [CrossRef]
[PubMed]
60. Du, X.-L.; Edelstein, D.; Rossetti, L.; Fantus, I.G.; Goldberg, H.; Ziyadeh, F.; Wu, J.; Brownlee, M.
Hyperglycemia-induced mitochondrial superoxide overproduction activates the hexosamine pathway
and induces plasminogen activator inhibitor-1 expression by increasing Sp1 glycosylation. Proc. Natl.
Acad. Sci. USA 2000, 97, 12222–12226. [CrossRef] [PubMed]
61. Gyorgy, P.; Norris, R.F.; Rose, C.S. Bifidus factor. I. A variant of Lactobacillus bifidus requiring a special growth
factor. Arch. Biochem. Biophys. 1954, 48, 193–201. [CrossRef]
62. Ward, R.E.; Ninonuevo, M.; Mills, D.A.; Lebrilla, C.B.; German, J.B. In vitro fermentation of breast milk
oligosaccharides by Bifidobacterium infantis and Lactobacillus gasseri. Appl. Environ. Microbiol. 2006, 72,
4497–4499. [CrossRef] [PubMed]
63. Ward, R.E.; Ninonuevo, M.; Mills, D.A.; Lebrilla, C.B.; German, J.B. In vitro fermentability of human milk
oligosaccharides by several strains of bifidobacteria. Mol. Nutr. Food Res. 2007, 51, 1398–1405. [CrossRef]
[PubMed]
64. Garrido, D.; Ruiz-Moyano, S.; Jimenez-Espinoza, R.; Eom, H.J.; Block, D.E.; Mills, D.A. Utilization of
galactooligosaccharides by Bifidobacterium longum subsp. infantis isolates. Food Microbiol. 2013, 33, 262–270.
[CrossRef] [PubMed]
65. LoCascio, R.; Ninonuevo, M.; Freeman, S.; Sela, D.; Grimm, R.; Lebrilla, C.B.; Mills, D.A.; German, J.B.
Glycoprofiling of bifidobacterial consumption of human milk oligosaccharides demonstrates strain specific,
preferential consumption of small chain glycans secreted in early human lactation. J. Agric. Food Chem. 2007,
55, 8914–8919. [CrossRef] [PubMed]
66. Marcobal, A.; Barboza, M.; Sonnenburg, E.D.; Pudlo, N.; Martens, E.C.; Desai, P.; Lebrilla, C.B.; Weimer, B.C.;
Mills, D.A.; German, J.B.; Sonnenburg, JL. Bacteroides in the infant gut consume milk oligosaccharides via
mucus-utilization pathways. Cell Host Microbe 2011, 10, 507–514. [CrossRef] [PubMed]
67. Ruiz-Moyano, S.; Totten, S.M.; Garrido, D.A.; Smilowitz, J.T.; German, J.B.; Lebrilla, C.B.; Mills, D.A. Variation
in consumption of human milk oligosaccharides by infant gut-associated strains of Bifidobacterium breve.
Appl. Environ. Microbiol. 2013, 79, 6040–6049. [CrossRef] [PubMed]
68. Sela, D.A.; Garrido, D.; Lerno, L.; Wu, S.; Tan, K.; Eom, H.-J.; Joachimiak, A.; Lebrilla, C.B.; Mills, D.A.
Bifidobacterium longum subsp. infantis ATCC 15697 α-fucosidases are active on fucosylated human milk
oligosaccharides. Appl. Environ. Microbiol. 2012, 78, 795–803. [CrossRef] [PubMed]
69. Sela, D.A.; Li, Y.; Lerno, L.; Wu, S.; Marcobal, A.M.; German, J.B.; Chen, X.; Lebrilla, C.B.; Mills, D.A.
An infant-associated bacterial commensal utilizes breast milk sialyloligosaccharides. J. Biol. Chem. 2011, 286,
11909–11918. [CrossRef] [PubMed]
70. Kitaoka, M. Bifidobacterial enzymes involved in the metabolism of human milk oligosaccharides. Adv. Nutr.
2012, 3, 422S–429S. [CrossRef] [PubMed]
71. James, K.; Motherway, M.O.; Penno, C.; O’Brien, R.L.; van Sinderen, D. Bifidobacterium breve UCC2003
employs multiple transcriptional regulators to control metabolism of particular human milk oligosaccharides.
Appl. Environ. Microbiol. 2018, 10, 278–279. [CrossRef] [PubMed]
72. Rudloff, S.; Pohlentz, G.; Diekmann, L.; Egge, H.; Kunz, C. Urinary excretion of lactose and oligosaccharides
in preterm infants fed human milk or infant formula. Acta Paediatr. 1996, 85, 598–603. [CrossRef] [PubMed]
73. Rudloff, S.; Pohlentz, G.; Borsch, C.; Lentze, M.J.; Kunz, C. Urinary excretion of in vivo 13 C-labelled milk
oligosaccharides in breastfed infants. Br. J. Nutr. 2012, 107, 957–963. [CrossRef] [PubMed]
74. Gnoth, M.J.; Rudloff, S.; Kunz, C.; Kinne, R.K. Investigations of the in vitro transport of human milk
oligosaccharides by a Caco-2 monolayer using a novel high performance liquid chromatography-mass
spectrometry technique. J. Biol Chem. 2001, 276, 34363–34370. [CrossRef] [PubMed]
75. Lane, J.A.; O’Callaghan, J.; Carrington, S.D.; Hickey, R.M. Transcriptional response of HT-29 intestinal
epithelial cells to human and bovine milk oligosaccharides. Br. J. Nutr. 2013, 110, 2127–2137. [CrossRef]
[PubMed]
76. He, Y.; Liu, S.; Kling, D.E.; Leone, S.; Lawlor, N.T.; Huang, Y.; Feinberg, S.B.; Hill, D.R.; Newburg, D.S.
The human milk oligosaccharide 2 -fucosyllactose modulates CD14 expression in human enterocytes, thereby
attenuating LPS-induced inflammation. Gut 2016, 65, 33–46. [CrossRef] [PubMed]
243
Nutrients 2018, 10, 1038
77. Bode, L.; Kunz, C.; Muhly-Reinholz, M.; Mayer, K.; Seeger, W.; Rudloff, S. Inhibition of monocyte, lymphocyte,
and neutrophil adhesion to endothelial cells by human milk oligosaccharides. Thromb. Haemost. 2004, 92,
1402–1410. [CrossRef] [PubMed]
78. Lin, A.E.; Autran, C.A.; Szyszka, A.; Escajadillo, T.; Huang, M.; Godula, K.; Prudden, A.R.; Boons, G.J.;
Lewis, A.L.; Doran, K.S.; et al. Human milk oligosaccharides inhibit growth of group B Streptococcus.
J. Biol. Chem. 2017, 292, 11243–11249. [CrossRef] [PubMed]
79. Yu, Z.T.; Nanthakumar, N.N.; Newburg, D.S. The human milk oligosaccharide 2 -fucosyllactose quenches
Campylobacter jejuni-induced inflammation in human epithelial cells HEp-2 and HT-29 and in mouse intestinal
mucosa. J. Nutr. 2016, 146, 1980–1990. [CrossRef] [PubMed]
80. Weichert, S.; Koromyslova, A.; Singh, B.K.; Hansman, S.; Jennewein, S.; Schroten, H.; Hansman, G.S.
Structural basis for Norovirus inhibition by human milk oligosaccharides. J. Virol. 2016, 90, 4843–4848.
[CrossRef] [PubMed]
81. Duska-McEwen, G.; Senft, A.P.; Ruetschilling, T.L.; Barrett, E.G.; Buck, R.H. Human milk oligosaccharides
enhance innate immunity to respiratory syncytial virus and influenza in vitro. Food Nutr. Sci. 2014, 5,
1387–1398. [CrossRef]
82. Hanisch, F.G.; Hansman, G.S.; Morozov, V.; Kunz, C.; Schroten, H. Avidity of α-fucose on human milk
oligosaccharides and blood group-unrelated oligo/polyfucoses is essential for potent norovirus-binding
targets. J. Biol. Chem. 2018, 293, 11955–11965. [CrossRef] [PubMed]
83. Li, M.; Monaco, M.H.; Wang, M.; Comstock, S.S.; Kuhlenschmidt, T.B.; Fahey, G.C., Jr.; Miller, M.J.;
Kuhlenschmidt, M.S.; Donovan, S.M. Human milk oligosaccharides shorten rotavirus-induced diarrhea and
modulate piglet mucosal immunity and colonic microbiota. ISME J. 2014, 8, 1609–1620. [CrossRef] [PubMed]
84. Comstock, S.S.; Li, M.; Wang, M.; Monaco, M.H.; Kuhlenschmidt, T.B.; Kuhlenschmidt, M.S.; Donovan, SM.
Dietary human milk oligosaccharides but not prebiotic oligosaccharides increase circulating natural killer
cell and mesenteric lymph node memory t cell populations in noninfected and rotavirus-infected neonatal
piglets. J. Nutr. 2017, 147, 1041–1047. [CrossRef] [PubMed]
85. Xiao, L.; Leusink-Muis, T.; Kettelarij, N.; van Ark, I.; Blijenberg, B.; Hesen, N.A.; Stahl, B.; Overbeek, S.A.;
Garssen, J.; Folkerts, G.; et al. Human milk oligosaccharide 2 -fucosyllactose improves innate and adaptive
immunity in an influenza-specific murine vaccination model. Front. Immunol. 2018, 9, 452. [CrossRef]
[PubMed]
86. Lewis, Z.T.; Totten, S.M.; Smilowitz, J.T.; Popovic, M.; Parker, E.; Lemay, D.G.; Van Tassell, M.L.; Miller, M.J.;
Jin, Y.S.; German, J.B.; et al. Maternal fucosyltransferase 2 status affects the gut bifidobacterial communities
of breastfed infants. Microbiome 2015, 3, 13. [CrossRef] [PubMed]
87. Ashida, H.; Miyake, A.; Kiyohara, M.; Wada, J.; Yoshida, E.; Kumagai, H.; Katayama, T.; Yamamoto, K.
Two distinct alpha-L-fucosidases from Bifidobacterium bifidum are essential for the utilization of fucosylated
milk oligosaccharides and glycoconjugates. Glycobiology 2009, 19, 1010–1017. [CrossRef] [PubMed]
88. Smith-Brown, P.; Morrison, M.; Krause, L.; Davies, P.S. Mothers secretor status affects development of
childrens microbiota composition and function: A pilot study. PLoS ONE 2016, 11, e0161211. [CrossRef]
[PubMed]
89. Sprenger, N.; Odenwald, H.; Kukkonen, A.K.; Kuitunen, M.; Savilahti, E.; Kunz, C. FUT2-dependent breast
milk oligosaccharides and allergy at 2 and 5 years of age in infants with high hereditary allergy risk.
Eur. J. Nutr. 2017, 56, 1293–1301. [CrossRef] [PubMed]
90. Seppo, A.E.; Autran, C.A.; Bode, L.; Järvinen, K.M. Human milk oligosaccharides and development of cow’s
milk allergy in infants. J. Allergy Clin. Immunol. 2017, 139, 708.e5–711.e5. [CrossRef] [PubMed]
91. Kuhn, L.; Kim, H.Y.; Hsiao, L.; Nissan, C.; Kankasa, C.; Mwiya, M.; Thea, D.M.; Aldrovandi, G.M.; Bode, L.
Oligosaccharide composition of breast milk influences survival of uninfected children born to HIV-infected
mothers in Lusaka, Zambia. J. Nutr. 2015, 145, 66–72. [CrossRef] [PubMed]
92. Morrow, A.L.; Ruiz-Palacios, G.M.; Altaye, M.; Jiang, X.; Guerrero, M.L.; Meinzen-Derr, J.K.; Farkas, T.;
Chaturvedi, P.; Pickering, L.K.; Newburg, D.S. Human milk oligosaccharides are associated with protection
against diarrhea in breast-fed infants. J. Pediatr. 2004, 145, 297–303. [CrossRef] [PubMed]
93. Marriage, B.; Buck, R.; Goehring, K.; Oliver, J.; Williams, J. Infants fed a lower caloric formula with 2 FL show
growth and 2 FL uptake like breast-fed infants. J. Pediatr. Gastroenterol. Nutr. 2015, 61, 649–658. [CrossRef]
[PubMed]
244
Nutrients 2018, 10, 1038
94. Goehring, K.; Marriage, B.; Oliver, J.; Wilder, J.; Barrett, E.; Buck, R. Similar to those who are breastfed,
infants fed a formula containing 2 -Fucosyllactose have lower inflammatory cytokines in a randomized
controlled trial. J. Nutr. 2016, 146, 2559–2566. [CrossRef] [PubMed]
95. Kajzer, J.; Oliver, J.; Marriage, B.F. Gastrointestinal tolerance of formula supplemented with oligosaccharides.
FASEB J. 2016, 30, 671.
96. Puccio, G.; Alliet, P.; Cajozzo, C.; Janssens, E.; Corsello, G.; Sprenger, N.; Wernimont, S.; Egli, D.; Gosoniu, L.;
Steenhout, P.L. Effects of infant formula with human milk oligosaccharides on growth and morbidity:
A randomized multicenter trial. J. Pediatr. Gastroenterol. Nutr. 2017, 64, 624–631. [CrossRef] [PubMed]
97. Steenhout, P.; Sperisen, P.; Martin, F.-P.; Sprenger, N.; Wernimont, S.; Pecquet, S.; Berger, B. Term infant
formula supplemented with human milk oligosaccharides (2 -fucosyllactose and lacto-N-neotetraose) shifts
stool microbiota and metabolic signatures closer to that of breastfed infants. FASEB J. 2016, 30, 275–277.
98. Berger, B.; Grathwohl, D.; Alliet, P.; Puccio, G.; Steenhout, P.; Sprenger, N. Stool microbiota in term infants
fed formula supplemented with human milk oligosaccharides and reduced likelihood of antibiotic use.
J. Pediatr. Gastroenterol. Nutr. 2016, 63, S407.
© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
245
nutrients
Review
Zinc and Sepsis
Wiebke Alker 1,2 and Hajo Haase 1,2, *
1 Department of Food Chemistry and Toxicology, Berlin Institute of Technology, 13355 Berlin, Germany;
alker@tu-berlin.de
2 TraceAge—DFG Research Unit on Interactions of Essential Trace Elements in Healthy and Diseased Elderly,
Potsdam-Berlin-Jena, Germany
* Correspondence: Haase@tu-berlin.de; Tel.: +49-30-314-727-01
1. Introduction
Zinc is of fundamental importance for the immune system and is involved in different pathologies.
In recent years, indications have appeared that zinc homeostasis might be an important factor
during sepsis. The following review focuses on the alterations of zinc homeostasis during sepsis
and possible physiological functions of this process. It further discusses potential risks and benefits of
zinc supplementation as well as a possible approach for using serum zinc as a biomarker for sepsis.
1.1. Sepsis
The term “sepsis” in relation to a disease has already been used by Hippocrates, but to this day it
remains a challenge to compile a definition comprising its complexity [1]. This results from the fact
that sepsis is rather a syndrome than an illness, showing a not yet fully elucidated pathobiology, and
with uniform diagnostic tests still lacking [2]. Sepsis is responsible for about 6 million deaths per year,
making it a critical illness and one of the major causes of mortality worldwide [3,4]. Its epidemiological
burden is assumed to be much higher in low- and middle-income countries and the mortality rate is
affected by the global national income [3,5].
There is an urgent need for an easily understandable definition in order to establish public
awareness, as well as for improved and uniform diagnostic guidelines for an early recognition of
sepsis [2,3]. In the past, different task forces have approached these issues [6,7]. A recent consensus
defined sepsis as a “life-threatening organ dysfunction caused by a dysregulated host-response to
infection” (Sepsis-3) [2]. To diagnose organ dysfunction in the clinical setting, Singer et al. recommend
the Sequential Organ Failure Assessment (SOFA) score. It includes parameters to evaluate the functions
of respiration, the liver, the cardiovascular system, the central nervous system, the kidneys, and
coagulation. An elevation of the total SOFA score of 2 points or more indicates organ dysfunction [2,8].
Sepsis is initiated by an infection [2]. The pathogen triggers an immune response, comprising
pro-inflammatory mechanisms to defeat the pathogen and regenerate the affected tissue, as well
as subsequent anti-inflammatory mechanisms to counteract the pro-inflammatory actions in order
to limit collateral damage in healthy tissue [9,10]. A dysregulation of this immune response, as
it appears during sepsis, leads to an over-reaction of the immune system, which can affect both
mechanisms described. Hyper-inflammation in the form of a systemic inflammatory response
syndrome (SIRS) can lead to a damage of the host’s own tissue. Immune-suppression, also known
as compensatory anti-inflammatory response syndrome (CARS), leaves the host more vulnerable
to secondary infections [2,11,12]. A wealth of literature is provided about sepsis and its symptoms,
diagnostics, and possible medical treatment approaches (e.g., [10,11,13,14]), to which the reader is
referred for more detailed information on these aspects of sepsis.
1.2. Zinc
Zinc is an essential trace element [15,16]. In the body it functions, for example, as a co-factor for a
high number of enzymes or as a structural element for a variety of proteins [17]. Zinc deficiency can
result in growth retardation, dermatitis, and hypogonadism, or symptoms such as delayed wound
healing, thymic atrophy or lymphopenia, and high incidence of infection; the latter points are due
to its particular importance for the immune system [18–20]. Consequently, zinc deficiency results
in multiple immunological changes, including what seems to be a shift toward a predominantly
innate immune response when the availability of zinc is limited [21]. One particularly important
effect of zinc is a modulation of the production of inflammatory cytokines [22]. Moreover, zinc is
crucial for the functioning of virtually all immune cells. For example, the differentiation of immature
T-cells depends on zinc, because thymulin, a hormone involved in T-cell differentiation, depends
on zinc as a co-factor [23,24]. In addition, the maturation of T-cells is influenced by their zinc
status. On the one hand a deficiency results in altered ratios of Th1- and Th2-cells, an increased
apoptosis-rate of immature T-cells, and consequently a decrease in T-cells in total [21,25–27]. On the
other hand, zinc supplementation has also been shown to promote regulatory T-cell development
and to suppress the maturation of Th17-cells, therefore having an inhibitory effect on Th17-mediated
autoimmune-diseases [28–30].
On the molecular level, some functions of zinc have been linked to its role as a second messenger
in immune cells. It has been shown that alterations in the intracellular free zinc-concentration function
as a “zinc signal”. Such a change in the intracellular free zinc concentration is induced by the binding
of various ligands to their respective receptors, such as lipopolysaccharide (LPS) to Toll-like receptor
4 (TLR-4), or the corresponding antigens to immunoglobulin E when it is present on the high-affinity
immunoglobulin E-receptor (FcεRI). Different kinds of immune cells vary in their expression of
receptors that utilize zinc; consequently zinc signals mediate diverse events, for example, formation
of pro-inflammatory cytokines by monocytes [31], presentation of major histocompatibility complex
(MHC) class II molecules at the surface of dendritic cells [32], formation of neutrophil extracellular
traps by neutrophil granulocytes [33], or proliferation of T-cells [34].
The essentiality of zinc for the immune system has been known since the 1960s and the
corresponding mechanistic knowledge has been expanding ever since. Its importance for the immune
system is based on various different mechanisms, each in its own way essential to ensure the
functionality of the immune system and the accurate processes of immune response, especially for
inflammatory processes. As a complete summary would exceed the scope of this article, the reader
is referred to recent review articles on the subject of zinc and immunity for more comprehensive
information [35,36] as well as to a recent review on the protective role of zinc during sepsis [37].
247
Nutrients 2018, 10, 976
248
Nutrients 2018, 10, 976
as to whether differences in gene expression in pediatric septic shock survivors and non-survivors
can be observed. Regarding zinc homeostasis, two isoforms of MT have been identified that showed
an increased expression in non-survivors compared to survivors. In addition, non-survivors had a
significantly lower serum zinc concentration compared to survivors. Considering the zinc-binding
properties of MT, Wong and colleagues interpret these results to indirectly imply that increased MT
expression in non survivors might affect zinc homeostasis and thereby serum zinc concentration [53].
Taken together, different approaches imply a contribution of zinc and its altered homeostasis to
the pathobiology of sepsis.
Figure 1. Possible functions of zinc in sepsis. During the APR of sepsis zinc is redistributed from serum
to liver. This process results in decreased serum zinc concentration and increased liver zinc. The altered
zinc concentrations seem to serve different functions and to be a part of the host’s defense against
pathogens. APR: acute phase reaction; IL: interleukin; TNF: tumor necrosis factor; MT: metallothionein;
APP: acute phase proteins.
Research on zinc homeostasis in the context of sepsis delivered a variety of explanations for the
beneficial effects of a redistribution of zinc. The respective studies are discussed below.
One of the main effects of the redistribution of zinc is an accumulation of zinc in the liver.
Hence, it seems as is if a higher liver zinc level might benefit the host during infection. Among other
things, the APR is not only characterized by the previously mentioned redistribution of zinc, but also
by production of acute phase proteins (APP) and the release of cytokines [56,57]. Zinc serves as an
249
Nutrients 2018, 10, 976
important structural element for many proteins and is required by enzymes involved in transcription
and translation. Therefore, the higher synthesis rate of APP in the liver could cause an increased
requirement for zinc during APR [56–58]. With respect to cytokine production a knockout (k.o.) of
ZIP14 in mice, a transporter important for the regulation of zinc homeostasis in hepatocytes, showed
lower mRNA expression of TNF-α, IL-6, IL-1β, and IL-10 in the liver compared to wild-type (w.t.)
mice after induction of sepsis in a murine model. Simultaneously, plasma levels of TNF-α, IL-6,
and IL-10 were significantly higher in k.o. than in w.t. mice. The results indicate a disadvantage of
the ZIP14 k.o. mice during sepsis based on increased markers of inflammation and an influence of
zinc, transported by ZIP14, on the production of cytokines during APR. However this observation is
surprising, because a decrease in mRNA expression would be expected to result in lower cytokine
levels. Possible explanations would be elevated cytokine expression elsewhere in the body, or an
impact on mechanisms mediating the expression of antagonists of the pro-inflammatory cytokines [50].
Other studies suggest that the altered zinc supply during endotoxemia has a major influence
on energy production in the liver. Injection of LPS caused an increase in hepatic zinc and MT in w.t.
mice, whereas zinc levels stayed unchanged in MT k.o. mice. At the same time the liver glucose of the
former stayed unchanged while the levels in MT k.o. mice decreased significantly. These results imply
a lack of hepatic gluconeogenesis in the MT KO mice and a role for MT, and most likely also for zinc,
in maintaining glycaemia after induction of an infection [59].
A protective role of zinc for the liver has also been suggested. Using murine models it was shown
that after injection of endotoxin, zinc-deficient nutrition resulted in enhanced lipid peroxidation in
the liver compared to the zinc adequate group [60]. Another study showed that zinc pre-treatment of
mice resulted in an increased intracellular availability of zinc in liver cells and was accompanied
by decreased accumulation of superoxide and necrotic cell death in the liver after injection of
LPS [61]. Both experimental observations support a protective role of zinc in the liver during infection.
Interestingly, after injection of endotoxin, Sakagouchi et al. saw an increase in MT only in the zinc
adequate group, but not in zinc-deficient animals, and therefore suggested a relation between zinc
concentration, endotoxin-induced MT and lipid peroxidation [60]. In contrast, Zhou et al. observed the
protective effects of zinc pre-treatment on liver cells in w.t. mice as well as MT k.o. mice, leading them
to propose an effect of zinc independent of MT [61]. These results are not necessarily contradictory,
but could imply that the protective effect of zinc on the liver could work in more than one way.
Further studies on this topic would be useful, since organ dysfunction is one of the hallmarks of
sepsis and a better understanding of its mechanisms is the basis of a possible prevention. In summary,
studies show multiple and diverse functions of zinc in the liver during the onset of sepsis, suggesting
a physiological basis for the accumulation of zinc.
The redistribution of zinc and accumulation in the liver is accompanied by a decrease in serum
zinc concentration. With regard to the host’s defense against pathogens, this effect might have some
benefits as well. One protective mechanism of the host is referred to as nutritional immunity. Pathogens,
just like all living organisms, require transition metals for their survival. The host’s strategy is to restrict
the pathogens’ access to essential transition metals, for example by lowering their concentrations in
the serum or secretion of metal ion binding proteins. This process is not limited to zinc but has been
described for other micronutrients, such as iron or manganese [62].
A decrease in the serum zinc concentration has also been shown to influence the respective
number and maturation of immune cells. Therefore, the alteration of serum zinc during the APR might
function as a signal. Using a murine model of zinc deficiency, a downregulation of lymphopoiesis
and upregulation of myelopoiesis was found [19,21,27,63]. In line with this observation, a decrease
of intracellular zinc occurred as a result of homeostatic changes during monocytic differentiation of
HL60 cells. Moreover, experiments simulating zinc deficiency showed that lower zinc levels promoted
the development of HL-60 cells along the myeloid lineage into functionally mature macrophages [64].
The immune cells that benefit from a decline of serum zinc are part of the innate immune system.
They represent the first line of host defense and provide a faster response than the cells of the adaptive
250
Nutrients 2018, 10, 976
immune system. Fraker and King proposed the hypothesis of a “reprogramming of the immune
system” during zinc deficiency in the form of a shift from adaptive immunity to predominantly innate
immunity [65]. Here, limited resources would be directed toward an immediate defense on the expense
of long-term protection. In the context of sepsis, the reduction of serum zinc as a promoting signal for
the innate immune system might be an attempt to focus the defense mechanisms toward a fast innate
immune reaction in the face of a potentially overwhelming infection.
The various aspects by which zinc homeostasis seems to be involved in the body’s defense
against pathogens suggest that the redistribution of zinc in the course of sepsis could serve multiple
physiological purposes. However, further research will be required to evaluate the physiological
significance of the different processes, thereby widening our understanding of the zinc-dependent
endogenous defense mechanisms. This knowledge is required to develop medical approaches in order
to support the host’s body and its defense during sepsis.
3. Zinc Supplementation
The correlation between low serum zinc concentrations and a higher mortality rate or chance of
recurrence raises the question, if supplementation of zinc might be a treatment option to improve the
outcomes of sepsis. Table 1 gives an overview of zinc supplementation studies in the context of sepsis in
humans. In these studies zinc supplementation took place after the onset of sepsis. Some of them show
a beneficial effect of zinc in form of a lower mortality rate and a better neurological development of
neonates [78–80]. However, it was also observed that supplementation did not result in any significant
differences between the zinc group and the control group [81] or even showed a harmful effect [82].
251
Nutrients 2018, 10, 976
The published animal studies mostly examined the effects of prophylactic zinc supplementation
(Table 2). In several of them the supplementation of zinc prior to induction of sepsis showed beneficial
effects, such as improved survival, lower serum concentrations of pro-inflammatory cytokines, lower
bacterial burden or improved pulmonary function compared to the control-group [50,83–85]. However,
for the animal studies the results are also not consistent and a missing effect of zinc supplementation is
reported [86] as well as a harmful outcome in a case where zinc was applied during the acute phase [87].
The negative results of zinc supplementation may be explained by the fact that the reduction of zinc
levels during sepsis may occur for a reason, such as the above-mentioned physiological functions.
High doses of zinc were shown to have a pro-inflammatory effect [88] and might therefore aggravate
inflammation, or zinc supplementation could potentially interfere with nutritional immunity, one of
the endogenous defense mechanisms based on a shortage of zinc in the serum [62].
Another important factor is the bioavailability of serum zinc. Albumin is the major zinc-buffering
protein [89] as well as a negative APP [56,90]. Its concentration decreases during sepsis, leading to a
decreased zinc-binding capacity of the serum [42]. As a consequence, supplementing sepsis patients
until the “normal” total serum zinc concentration is restored would result in a supraphysiological
concentration of free, and thereby bioavailable, zinc in these patients.
Correcting zinc deficiency prior to sepsis is certainly beneficial, but difficult to realize, as in
most cases sepsis cannot be predicted. Moderate supplementation during sepsis might, in some
cases, turn out to be helpful, as well. This may be particularly so in patients with pre-existing zinc
deficiency that is so pronounced that the liver cannot accumulate sufficient amounts to exert the
abovementioned protective effects of zinc. However, as the reduction of serum zinc seems to be a
necessary physiological process, in these cases extreme care needs to be taken in order not to exceed
the zinc-binding capacity of the serum to avoid negative effects, such as counteracting nutritional
immunity or aggravating inflammation.
The endogenous processes, which are supposed to be affected by the supplementation of zinc,
are quite complex and fine-tuned. As illustrated by the divergence of the study results, effective zinc
supplementation has to be just as elaborate with regard to timing as well as dosage in order to achieve
optimal results.
252
Table 1. Zinc supplementation studies in septic patients.
Observation Time
Study Population Intervention/Zn-Supply Results (Zinc Group vs. Control Group) Reference
Points
Zinc group *:
Antibiotic treatment, dose of
Nutrients 2018, 10, 976
253
tests BW Zn2+ per day) abnormalities is 70% less) at one month of age
Control group: • Similar duration of hospitalization
Antibiotic treatment
Zinc group *:
Antibiotic treatment, dose of • Significant increase in serum zinc
3 mg/kg BW zinc sulfate concentrations
Neonates with clinical
monohydrate twice a day for Measurement of blood • Significantly lower mortality rate (6.6%
manifestations of sepsis
10 days samples from BL and compared to 17.3%) [80]
who exhibited two
(corresponding to 2.1 mg/kg after 10 days • Better neurodevelopment (significantly better
positive screening tests
BW Zn2+ per day) Mental Development Quotient) at 12 month of
Control group: age
Antibiotic treatment
Table 1. Cont.
Observation Time
Study Population Intervention/Zn-Supply Results (Zinc Group vs. Control Group) Reference
Points
Zinc group *:
Antibiotic treatment, dose of
Nutrients 2018, 10, 976
254
on day 3 [82]
or catheter sepsis Zn2+ per day) temperatures from
• No difference in serum IL-6 and ceruloplasmin
Control group: patients’ bedside charts
Total parental nutrition, 0 mg from day 1, 2, 3
zinc sulfate for 3 days
* In all studies, zinc supplementation was started after the onset of sepsis.
Table 2. Zinc supplementation studies in animal models of sepsis. LPS: lipopolysaccharide.
Animals Sepsis Model Intervention/Zn-Supply Results (Zinc Group vs. Control Group) Reference
Zinc group:
Injection of 10 mg/kg BW zinc
• Significantly improved survival following sepsis at 72
gluconate every 24 h for
Nutrients 2018, 10, 976
h after induction
3 days prior to induction of
Intraperitoneal (i. p.) • Significantly lower myeloperoxidase activity in lung
sepsis, injection continued
fecal slurry injection; tissue (at 24 h)
every 24 h after induction of
Male mice sacrifice of mice at 24 h • Significantly lower bacterial burden in blood and
peritonitis [84]
(C57BL/6) to conduct assays or spleen (at 24 h)
(corresponding to 1.4 mg/kg
observed 72 h for • Significantly lower serum keratinocyte
BW Zn2+ per day)
survival study chemoattractant concentration (at 24 h)
Control group:
• No significant difference between serum concentration
Injection of equal volume of
of IL6, IL-1β, IL-10
saline at the same time points
as for the zinc group
• Significantly lower IL-6 mRNA expression in
hepatocytes
Zinc group: • Significantly lower TNF-α mRNA expression in
255
High-zinc diet (180 mg/kg) hepatic leukocytes
for 7 days prior to induction • Significantly lower S100A9 mRNA expression in white
Cecal ligation and
Male and female mice of sepsis blood cells
puncture; sacrifice of [50]
(C57BL/6) Control group: • Significantly lower serum concentrations of TNF-α,
mice at 24 h
Zinc-adequate diet (30 mg/kg) S100A8 and S100A9
for 7 days prior to induction • Significantly lower serum concentration of plasma
of sepsis alanine aminotransferase
• Significantly lower bacterial burden in blood
and spleen
Zinc group:
Injection of 10 mg/kg BW zinc • Significantly improved survival of following sepsis at
I. p. cecal-slurry gluconate once a day for 72 h after induction
injection and 3 days prior to induction of • Significantly lower myeloperoxidase activity in lung
Male and female juvenile measurement of blood sepsis tissue (at 12 h)
mice samples at 6 h and 12 h; (corresponding to 1.4 mg/kg • Significantly lower bacterial burden in peritoneal fluid [83]
(C57BL/6) mice were sacrificed at 6 BW Zn2+ per day) (at 12 h)
h or 12 h or observed for Control group: • Significantly lower serum concentrations of IL-2 (at 6
72 h for survival study Injection of equal volume of h, 12 h), IL-6 (at 6 h, 12 h), IL-1β (at 6 h), and
saline for 3 days prior to keratinocyte-derived chemokines (at 12 h)
induction of sepsis
Table 2. Cont.
Animals Sepsis Model Intervention/Zn-Supply Results (Zinc Group vs. Control Group) Reference
Zinc group:
Intravenous infusion of • Increased arterial and venous oxygen pressure
Infusion of 25 mg/kg BW
LPS and measurement of (reaching significance at 45 min or 210 min)
zinc-bis-(DL-hydrogenaspartate)
Nutrients 2018, 10, 976
256
(Deutsche Landrasse) infusion of LPS; pigs
BW Zn2+ per day) • Increase in mean hemoglobin (reaching significance at
were sacrificed at 60 min
Control group: 30 min)
and organs removed for
Infusion of saline 2 h prior to • Increase in IL-6 and TNF-α plasma concentrations
analysis
infusion of LPS (reaching significance at 0 min or 45 min)
• Significant higher weights of lungs, width of alveolar
septae and rate of paracentral liver necrosis
Zinc group:
Intravenous infusion of Infusion of LPS at 0 h, 5 h and
LPS and measurement of 12 h, infusion of 25 mg/kg BW
• Trend to higher arterial and venous oxygen pressure
parameters for a zinc-bis-(DL-hydrogenaspartate)
• Trend to higher arterial and venous oxygen saturation
duration of 1020 min, (corresponding to 5 mg/kg
Female farm pigs • No significant differences in intrapulmonary shunt
with infusion of zinc BW Zn2+ per day) [86]
(Deutsche Landrasse) • No significant differences in EVLW
from 600 to 720 min; pigs at 10 h during sepsis
• Different courses in IL-6 and TNF-α plasma
were sacrificed at the end Control group:
concentrations, at the end almost similar levels
of the study period and a Infusion of LPS at 0 h, 5 h and
necropsy carried out 12 h, infusion of saline at 10 h
during sepsis
Nutrients 2018, 10, 976
257
Nutrients 2018, 10, 976
5. Conclusions
Zinc is an essential trace element and has been shown to be crucial for ensuring an adequate
immune response. In the context of sepsis the host’s zinc homeostasis is altered. Various study results
imply some of the alterations to be part of the host’s defense mechanism against pathogens (Figure 1).
There are indications that a patient’s zinc supply and serum zinc concentration is associated with
severity, outcome, and recurrence of sepsis. Zinc seems to have potential to be used as a biomarker
or even as a starting point for a therapeutic approach. Further research is required to broaden the
understanding of zinc homeostasis during sepsis and the underlying mechanisms as well as to evaluate
the possible clinical applicability of this knowledge.
Funding: This work was funded by a grant from Deutsche Forschungsgemeinschaft (HA 4318/4-1) within
the TraceAge—DFG Research Unit on Interactions of essential trace elements in healthy and diseased elderly,
Potsdam-Berlin-Jena, FOR 2558/1.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Funk, D.J.; Parrillo, J.E.; Kumar, A. Sepsis and septic shock: A history. Crit. Care Clin. 2009, 25, 83–101.
[CrossRef] [PubMed]
2. Singer, M.; Deutschman, C.S.; Seymour, C.W.; Shankar-Hari, M.; Annane, D.; Bauer, M.; Bellomo, R.;
Bernard, G.R.; Chiche, J.-D.; Coopersmith, C.M.; et al. The third international consensus definitions for sepsis
and septic shock (Sepsis-3). JAMA 2016, 315, 801–810. [CrossRef] [PubMed]
3. WHO (The World Health Organization). WHA Resolution A70/13—Improving the Prevention, Diagnosis and
Clinical Management of Sepsis; WHO: Geneva, Switzerland, 2017.
4. WHO (The World Health Organization). WHO Secretariat Report A70/13—Improving the Prevention, Diagnosis
and Clinical Management of Sepsis; WHO: Geneva, Switzerland, 2017.
5. Vincent, J.-L.; Marshall, J.C.; Ñamendys-Silva, S.A.; François, B.; Martin-Loeches, I.; Lipman, J.; Reinhart, K.;
Antonelli, M.; Pickkers, P.; Njimi, H.; et al. Assessment of the worldwide burden of critical illness:
The Intensive Care Over Nations (ICON) audit. Lancet Respir. Med. 2014, 2, 380–386. [CrossRef]
6. Bone, R.C.; Balk, R.A.; Cerra, F.B.; Dellinger, R.P.; Fein, A.M.; Knaus, W.A.; Schein, R.M.H.; Sibbald, W.J.
Definitions for Sepsis and Organ Failure and Guidelines for the Use of Innovative Therapies in Sepsis. Chest
1992, 101, 1644–1655. [CrossRef] [PubMed]
7. Levy, M.M.; Fink, M.P.; Marshall, J.C.; Abraham, E.; Angus, D.; Cook, D.; Cohen, J.; Opal, S.M.; Vincent, J.;
Ramsay, G. 2001 SCCM/ESICM/ACCP/ATS/SIS international sepsis definitions conference. Crit. Care Med.
2003, 31, 1250–1256. [CrossRef] [PubMed]
8. Vincent, J.-L.; Moreno, R.; Takala, J.; Willatts, S.; Mendonça, A.D.; Bruining, H.; Reinhart, C.K.; Suter, P.M.;
Thijs, L.G. The SOFA (Sepsis-related Organ Failure Assessment) score to describe organ dysfunction/failure.
Intensive Care Med. 1996, 22, 707–710. [CrossRef] [PubMed]
9. Takeuchi, O.; Akira, S. Pattern recognition receptors and inflammation. Cell 2010, 140, 805–820. [CrossRef]
[PubMed]
10. Van der Poll, T.; Opal, S.M. Host–pathogen interactions in sepsis. Lancet Infect. Dis. 2008, 8, 32–43. [CrossRef]
11. Angus, D.C.; van der Poll, T. Severe sepsis and septic shock. N. Engl. J. Med. 2013, 369, 840–851. [CrossRef]
[PubMed]
12. Bone, R.C.; Grodzin, C.J.; Balk, R.A. Sepsis: A New hypothesis for pathogenesis of the disease process.
CHEST 1997, 112, 235–243. [CrossRef] [PubMed]
13. Chong, J.; Dumont, T.; Francis-frank, L.; Balaan, M. Sepsis and septic shock: A review. Crit. Care Nurs. Q.
2015, 38, 111–120. [CrossRef] [PubMed]
14. Ward, N.S.; Casserly, B.; Ayala, A. The Compensatory Anti-inflammatory Response syndrome (CARS) in
critically ill patients. Clin. Chest Med. 2008, 29, 617–625. [CrossRef] [PubMed]
15. Prasad, A.S.; Halsted, J.A.; Nadimi, M. Syndrome of iron deficiency anemia, hepatosplenomegaly,
hypogonadism, dwarfism and geophagia. Am. J. Med. 1961, 31, 532–546. [CrossRef]
16. Prasad, A.S. Importance of zinc in human nutrition. Am. J. Clin. Nutr. 1967, 20, 648–652. [CrossRef]
[PubMed]
258
Nutrients 2018, 10, 976
17. Coleman, J.E. Zinc proteins: Enzymes, storage proteins, transcription factors, and replication proteins.
Annu. Rev. Biochem. 1992, 61, 897–946. [CrossRef] [PubMed]
18. Evans, G.W. Zinc and its deficiency diseases. Clin. Physiol. Biochem. 1986, 4, 94–98. [PubMed]
19. King, L.E.; Frentzel, J.W.; Mann, J.J.; Fraker, P.J. Chronic zinc deficiency in mice disrupted T cell lymphopoiesis
and erythropoiesis while B cell lymphopoiesis and myelopoiesis were maintained. J. Am. Coll. Nutr. 2005,
24, 494–502. [CrossRef] [PubMed]
20. Prasad, A.S.; Meftah, S.; Abdallah, J.; Kaplan, J.; Brewer, G.J.; Bach, J.F.; Dardenne, M. Serum thymulin in
human zinc deficiency. J. Clin. Investig. 1988, 82, 1202–1210. [CrossRef] [PubMed]
21. Fraker, P.J.; King, L.E. A distinct role for apoptosis in the changes in lymphopoiesis and myelopoiesis created
by deficiencies in zinc. FASEB J. 2001, 15, 2572–2578. [CrossRef] [PubMed]
22. Mayer, L.S.; Uciechowski, P.; Meyer, S.; Schwerdtle, T.; Rink, L.; Haase, H. Differential impact of zinc
deficiency on phagocytosis, oxidative burst, and production of pro-inflammatory cytokines by human
monocytes. Metallomics 2014, 6, 1288–1295. [CrossRef] [PubMed]
23. Dardenne, M.; Pléau, J.M.; Nabarra, B.; Lefrancier, P.; Derrien, M.; Choay, J.; Bach, J.F. Contribution of zinc
and other metals to the biological activity of the serum thymic factor. Proc. Natl. Acad. Sci. USA 1982,
79, 5370–5373. [CrossRef] [PubMed]
24. Incefy, G.S.; Mertelsmann, R.; Yata, K.; Dardenne, M.; Bach, J.F.; Good, R.A. Induction of differentiation
in human marrow T cell precursors by the synthetic serum thymic factor, FTS. Clin. Exp. Immunol. 1980,
40, 396–406. [PubMed]
25. Prasad, A.S. Effects of zinc deficiency on Th1 and Th2 cytokine shifts. J. Infect. Dis. 2000, 182, S62–S68.
[CrossRef] [PubMed]
26. Beck, F.W.; Prasad, A.S.; Kaplan, J.; Fitzgerald, J.T.; Brewer, G.J. Changes in cytokine production and T cell
subpopulations in experimentally induced zinc-deficient humans. Am. J. Physiol. Endocrinol. Metab. 1997,
272, E1002–E1007. [CrossRef] [PubMed]
27. King, L.E.; Osati-Ashtiani, F.; Fraker, P.J. Apoptosis plays a distinct role in the loss of precursor lymphocytes
during zinc deficiency in mice. J. Nutr. 2002, 132, 974–979. [CrossRef] [PubMed]
28. Rosenkranz, E.; Maywald, M.; Hilgers, R.-D.; Brieger, A.; Clarner, T.; Kipp, M.; Plümäkers, B.; Meyer, S.;
Schwerdtle, T.; Rink, L. Induction of regulatory T cells in Th1-/Th17-driven experimental autoimmune
encephalomyelitis by zinc administration. J. Nutr. Biochem. 2016, 29, 116–123. [CrossRef] [PubMed]
29. Lee, H.; Kim, B.; Choi, Y.H.; Hwang, Y.; Kim, D.H.; Cho, S.; Hong, S.J.; Lee, W. Inhibition of
interleukin-1β-mediated interleukin-1 receptor-associated kinase 4 phosphorylation by zinc leads to
repression of memory T helper type 17 response in humans. Immunology 2015, 146, 645–656. [CrossRef]
[PubMed]
30. Kitabayashi, C.; Fukada, T.; Kanamoto, M.; Ohashi, W.; Hojyo, S.; Atsumi, T.; Ueda, N.; Azuma, I.; Hirota, H.;
Murakami, M.; et al. Zinc suppresses Th17 development via inhibition of STAT3 activation. Int. Immunol.
2010, 22, 375–386. [CrossRef] [PubMed]
31. Haase, H.; Ober-Blöbaum, J.L.; Engelhardt, G.; Hebel, S.; Heit, A.; Heine, H.; Rink, L. Zinc signals are
essential for lipopolysaccharide-induced signal transduction in monocytes. J. Immunol. 2008, 181, 6491–6502.
[CrossRef] [PubMed]
32. Kitamura, H.; Morikawa, H.; Kamon, H.; Iguchi, M.; Hojyo, S.; Fukada, T.; Yamashita, S.; Kaisho, T.; Akira, S.;
Murakami, M.; et al. Toll-like receptor–mediated regulation of zinc homeostasis influences dendritic cell
function. Nat. Immunol. 2006, 7, 971. [CrossRef] [PubMed]
33. Hasan, R.; Rink, L.; Haase, H. Zinc signals in neutrophil granulocytes are required for the formation of
neutrophil extracellular traps. Innate Immun. 2013, 19, 253–264. [CrossRef] [PubMed]
34. Kaltenberg, J.; Plum, L.M.; Ober-Blöbaum, J.L.; Hönscheid, A.; Rink, L.; Haase, H. Zinc signals promote
IL-2-dependent proliferation of T cells. Eur. J. Immunol. 2010, 40, 1496–1503. [CrossRef] [PubMed]
35. Maares, M.; Haase, H. Zinc and immunity: An essential interrelation. Arch. Biochem. Biophys. 2016, 611, 58–65.
[CrossRef] [PubMed]
36. Wessels, I.; Maywald, M.; Rink, L. Zinc as a Gatekeeper of Immune Function. Nutrients 2017, 9, 1286.
[CrossRef] [PubMed]
37. Souffriau, J.; Libert, C. Mechanistic insights into the protective impact of zinc on sepsis. Cytokine Growth
Factor Rev. 2017. [CrossRef] [PubMed]
259
Nutrients 2018, 10, 976
38. Gaetke, L.M.; McClain, C.J.; Talwalkar, R.T.; Shedlofsky, S.I. Effects of endotoxin on zinc metabolism in
human volunteers. Am. J. Physiol. Endocrinol. Metab. 1997, 272, E952–E956. [CrossRef] [PubMed]
39. Besecker, B.Y.; Exline, M.C.; Hollyfield, J.; Phillips, G.; DiSilvestro, R.A.; Wewers, M.D.; Knoell, D.L. A
comparison of zinc metabolism, inflammation, and disease severity in critically ill infected and noninfected
adults early after intensive care unit admission123. Am. J. Clin. Nutr. 2011, 93, 1356–1364. [CrossRef]
[PubMed]
40. Mertens, K.; Lowes, D.A.; Webster, N.R.; Talib, J.; Hall, L.; Davies, M.J.; Beattie, J.H.; Galley, H.F. Low zinc
and selenium concentrations in sepsis are associated with oxidative damage and inflammation. Br. J. Anaesth.
2015, 114, 990–999. [CrossRef] [PubMed]
41. Hoeger, J.; Simon, T.-P.; Beeker, T.; Marx, G.; Haase, H.; Schuerholz, T. Persistent low serum zinc is associated
with recurrent sepsis in critically ill patients—A pilot study. PLoS ONE 2017, 12, e0176069. [CrossRef]
[PubMed]
42. Hoeger, J.; Simon, T.-P.; Doemming, S.; Thiele, C.; Marx, G.; Schuerholz, T.; Haase, H. Alterations in zinc
binding capacity, free zinc levels and total serum zinc in a porcine model of sepsis. BioMetals 2015, 28, 693–700.
[CrossRef] [PubMed]
43. Liuzzi, J.P.; Lichten, L.A.; Rivera, S.; Blanchard, R.K.; Aydemir, T.B.; Knutson, M.D.; Ganz, T.; Cousins, R.J.
Interleukin-6 regulates the zinc transporter Zip14 in liver and contributes to the hypozincemia of the
acute-phase response. Proc. Natl. Acad. Sci. USA 2005, 102, 6843–6848. [CrossRef] [PubMed]
44. Lichten, L.A.; Liuzzi, J.P.; Cousins, R.J. Interleukin-1β contributes via nitric oxide to the upregulation and
functional activity of the zinc transporter Zip14 (Slc39a14) in murine hepatocytes. Am. J. Physiol. Gastrointest.
Liver Physiol. 2009, 296, G860–G867. [CrossRef] [PubMed]
45. Sobocinski, P.Z.; Canterbury, W.J.; Mapes, C.A.; Dinterman, R.E. Involvement of hepatic metallothioneins
in hypozincemia associated with bacterial infection. Am. J. Physiol. Endocrinol. Metab. 1978, 234, E399.
[CrossRef] [PubMed]
46. Sobocinski, P.Z.; Canterbury, W.J. Hepatic metallothionein induction in inflammation. Ann. N. Y. Acad. Sci.
1982, 389, 354–367. [CrossRef]
47. Cousins, R.J.; Leinart, A.S. Tissue-specific regulation of zinc metabolism and metallothionein genes by
interleukin 1. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 1988, 2, 2884–2890. [CrossRef]
48. Coyle, P.; Philcox, J.C.; Carey, L.C.; Rofe, A.M. Metallothionein: The multipurpose protein. Cell. Mol. Life
Sci. CMLS 2002, 59, 627–647. [CrossRef] [PubMed]
49. Huber, K.L.; Cousins, R.J. Metallothionein expression in rat bone marrow is dependent on dietary zinc but
not dependent on interleukin-1 or interleukin-6. J. Nutr. 1993, 123, 642–648. [CrossRef] [PubMed]
50. Wessels, I.; Cousins, R.J. Zinc dyshomeostasis during polymicrobial sepsis in mice involves zinc transporter
Zip14 and can be overcome by zinc supplementation. Am. J. Physiol. Gastrointest. Liver Physiol. 2015,
309, G768–G778. [CrossRef] [PubMed]
51. Rech, M.; To, L.; Tovbin, A.; Smoot, T.; Mlynarek, M. Heavy metal in the intensive care unit: A review of
current literature on trace element supplementation in critically ill patients. Nutr. Clin. Pract. 2014, 29, 78–89.
[CrossRef] [PubMed]
52. Wong, H.R. Pediatric septic shock treatment: New clues from genomic profiling. Pharmacogenomics 2007,
8, 1287–1290. [CrossRef] [PubMed]
53. Wong, H.R.; Shanley, T.P.; Sakthivel, B.; Cvijanovich, N.; Lin, R.; Allen, G.L.; Thomas, N.J.; Doctor, A.;
Kalyanaraman, M.; Tofil, N.M.; et al. Genome-level expression profiles in pediatric septic shock indicate a
role for altered zinc homeostasis in poor outcome. Physiol. Genom. 2007, 30, 146–155. [CrossRef] [PubMed]
54. Shanley, T.P.; Wong, H.R. Molecular genetics in the pediatric intensive care unit. Crit. Care Clin. 2003,
19, 577–594. [CrossRef]
55. Cvijanovich, N.; Shanley, T.P.; Lin, R.; Allen, G.L.; Thomas, N.J.; Checchia, P.; Anas, N.; Freishtat, R.J.;
Monaco, M.; Odoms, K.; et al. Validating the genomic signature of pediatric septic shock. Physiol. Genom.
2008, 34, 127–134. [CrossRef] [PubMed]
56. Moshage, H. Cytokines and the hepatic acute phase response. J. Pathol. 1997, 181, 257–266. [CrossRef]
57. Baumann, H.; Gauldie, J. The acute phase response. Immunol. Today 1994, 15, 74–80. [CrossRef]
58. Florea, D.; Molina-López, J.; Hogstrand, C.; Lengyel, I.; de la Cruz, A.P.; Rodríguez-Elvira, M.; Planells, E.
Changes in zinc status and zinc transporters expression in whole blood of patients with Systemic
Inflammatory Response Syndrome (SIRS). J. Trace Elem. Med. Biol. 2017. [CrossRef] [PubMed]
260
Nutrients 2018, 10, 976
59. Rofe, A.M.; Philcox, J.C.; Coyle, P. Trace metal, acute phase and metabolic response to endotoxin in
metallothionein-null mice. Biochem. J. 1996, 314, 793–797. [CrossRef] [PubMed]
60. Sakaguchi, S.; Iizuka, Y.; Furusawa, S.; Ishikawa, M.; Satoh, S.; Takayanagi, M. Role of Zn2+ in oxidative
stress caused by endotoxin challenge. Eur. J. Pharmacol. 2002, 451, 309–316. [CrossRef]
61. Zhou, Z.; Wang, L.; Song, Z.; Saari, J.T.; McClain, C.J.; Kang, Y.J. Abrogation of nuclear factor-κB activation
is involved in zinc inhibition of lipopolysaccharide-induced tumor necrosis factor-α production and liver
injury. Am. J. Pathol. 2004, 164, 1547–1556. [CrossRef]
62. Hood, M.I.; Skaar, E.P. Nutritional immunity: Transition metals at the pathogen-host interface.
Nat. Rev. Microbiol. 2012, 10. [CrossRef] [PubMed]
63. King, L.E.; Osati-Ashtiani, F.; Fraker, P.J. Depletion of cells of the B lineage in the bone marrow of
zinc-deficient mice. Immunology 1995, 85, 69–73. [PubMed]
64. Dubben, S.; Hönscheid, A.; Winkler, K.; Rink, L.; Haase, H. Cellular zinc homeostasis is a regulator in
monocyte differentiation of HL-60 cells by 1α,25-dihydroxyvitamin D3. J. Leukoc. Biol. 2010, 87, 833–844.
[CrossRef] [PubMed]
65. Fraker, P.J.; King, L.E. Reprogramming of the immune system during zinc deficiency. Annu. Rev. Nutr. 2004,
24, 277–298. [CrossRef] [PubMed]
66. Prasad, A.S.; Beck, F.W.; Bao, B.; Fitzgerald, J.T.; Snell, D.C.; Steinberg, J.D.; Cardozo, L.J.
Zinc supplementation decreases incidence of infections in the elderly: Effect of zinc on generation of
cytokines and oxidative stress. Am. J. Clin. Nutr. 2007, 85, 837–844. [CrossRef] [PubMed]
67. Wessels, I.; Haase, H.; Engelhardt, G.; Rink, L.; Uciechowski, P. Zinc deficiency induces production of
the proinflammatory cytokines IL-1β and TNFα in promyeloid cells via epigenetic and redox-dependent
mechanisms. J. Nutr. Biochem. 2013, 24, 289–297. [CrossRef] [PubMed]
68. Oteiza, P.I.; Clegg, M.S.; Zago, M.P.; Keen, C.L. Zinc deficiency induces oxidative stress and AP-1 activation
in 3T3 cells. Free Radic. Biol. Med. 2000, 28, 1091–1099. [CrossRef]
69. Oteiza, P.I.; Olin, K.L.; Fraga, C.G.; Keen, C.L. Zinc deficiency causes oxidative damage to proteins, lipids
and DNA in rat testes. J. Nutr. 1995, 125, 823–829. [CrossRef] [PubMed]
70. Song, Y.; Chung, C.S.; Bruno, R.S.; Traber, M.G.; Brown, K.H.; King, J.C.; Ho, E. Dietary zinc restriction and
repletion affects DNA integrity in healthy men. Am. J. Clin. Nutr. 2009, 90, 321–328. [CrossRef] [PubMed]
71. Hotchkiss, R.S.; Osmon, S.B.; Chang, K.C.; Wagner, T.H.; Coopersmith, C.M.; Karl, I.E. Accelerated
lymphocyte death in sepsis occurs by both the death receptor and mitochondrial pathways. J. Immunol. 2005,
174, 5110–5118. [CrossRef] [PubMed]
72. Hotchkiss, R.S.; Swanson, P.E.; Freeman, B.D.; Tinsley, K.W.; Cobb, J.P.; Matuschak, G.M.; Buchman, T.G.;
Karl, I.E. Apoptotic cell death in patients with sepsis, shock, and multiple organ dysfunction. Crit. Care Med.
1999, 27, 1230. [CrossRef] [PubMed]
73. Heidecke, C.-D.; Hensler, T.; Weighardt, H.; Zantl, N.; Wagner, H.; Siewert, J.-R.; Holzmann, B. Selective
defects of T lymphocyte function in patients with lethal intraabdominal infection. Am. J. Surg. 1999,
178, 288–292. [CrossRef]
74. Andresen, M.; Regueira, T.; Bruhn, A.; Perez, D.; Strobel, P.; Dougnac, A.; Marshall, G.; Leighton, F.
Lipoperoxidation and protein oxidative damage exhibit different kinetics during septic shock.
Mediat. Inflamm. 2008, 2008. [CrossRef] [PubMed]
75. Kaymak, C.; Kadioglu, E.; Ozcagli, E.; Osmanoglu, G.; Izdes, S.; Agalar, C.; Basar, H.; Sardas, S.
Oxidative DNA damage and total antioxidant status in rats during experimental gram-negative sepsis.
Hum. Exp. Toxicol. 2008, 27, 485–491. [CrossRef] [PubMed]
76. Galley, H.F. Oxidative stress and mitochondrial dysfunction in sepsis. Br. J. Anaesth. 2011, 107, 57–64.
[CrossRef] [PubMed]
77. Cander, B.; Dundar, Z.D.; Gul, M.; Girisgin, S. Prognostic value of serum zinc levels in critically ill patients.
J. Crit. Care 2011, 26, 42–46. [CrossRef] [PubMed]
78. Newton, B.; Ballambattu, V.B.; Bosco, D.B.; Gopalakrishna, S.M.; Subash, C.P. Efficacy of
zinc supplementation on serum calprotectin, inflammatory cytokines and outcome in neonatal
sepsis—Arandomized controlled trial. J. Matern. Fetal Neonatal Med. 2017, 30, 1627–1631. [CrossRef]
79. Newton, B.; Bhat, B.V.; Bosco Dhas, B.; Mondal, N.; Gopalakrishna, S.M. Effect of zinc supplementation
on early outcome of neonatal sepsis—A randomized controlled trial. Indian J. Pediatr. 2016, 83, 289–293.
[CrossRef] [PubMed]
261
Nutrients 2018, 10, 976
80. Newton, B.; Bhat, B.V.; Bosco Dhas, B.; Christina, C.; Gopalakrishna, S.M.; Subhash Chandra, P. Short
term oral zinc supplementation among babies with neonatal sepsis for reducing mortality and improving
outcome—A double-blind randomized controlled trial. Indian J. Pediatr. 2018, 85, 5–9. [CrossRef]
81. Mehta, K.; Bhatta, N.K.; Majhi, S.; Shrivastava, M.K.; Singh, R.R. Oral zinc supplementation for reducing
mortality in probable neonatal sepsis: A double blind randomized placebo controlled trial. Indian Pediatr.
2013, 50, 390–393. [CrossRef] [PubMed]
82. Braunschweig, C.L.; Sowers, M.; Kovacevich, D.S.; Hill, G.M.; August, D.A. Parenteral zinc supplementation
in adult humans during the acute phase response increases the febrile response. J. Nutr. 1997, 127, 70–74.
[CrossRef] [PubMed]
83. Ganatra, H.A.; Varisco, B.M.; Harmon, K.; Lahni, P.; Opoka, A.; Wong, H.R. Zinc supplementation leads
to immune modulation and improved survival in a juvenile model of murine sepsis. Innate Immun. 2017,
23, 67–76. [CrossRef] [PubMed]
84. Nowak, J.E.; Harmon, K.; Caldwell, C.C.; Wong, H.R. Prophylactic zinc supplementation reduces bacterial
load and improves survival in a murine model of sepsis. Pediatr. Crit. Care Med. J. Soc. Crit. Care Med. World
Fed. Pediatr. Intensive Crit. Care Soc. 2012, 13, e323–e329. [CrossRef] [PubMed]
85. Krones, C.J.; Klosterhalfen, B.; Butz, N.; Hoelzl, F.; Junge, K.; Stumpf, M.; Peiper, C.; Klinge, U.;
Schumpelick, V. Effect of zinc pretreatment on pulmonary endothelial cells in vitro and pulmonary function
in a porcine model of endotoxemia. J. Surg. Res. 2005, 123, 251–256. [CrossRef] [PubMed]
86. Krones, C.J.; Klosterhalfen, B.; Anurov, M.; Stumpf, M.; Klinge, U.; Oettinger, A.P.; Schumpelick, V. Missing
effects of zinc in a porcine model of recurrent endotoxemia. BMC Surg. 2005, 5, 22. [CrossRef] [PubMed]
87. Krones, C.J.; Klosterhalfen, B.; Fackeldey, V.; Junge, K.; Rosch, R.; Schwab, R.; Stumpf, M.; Klinge, U.;
Schumpelick, V. Deleterious effect of zinc in a pig model of acute endotoxemia. J. Investig. Surg. 2004,
17, 249–256. [CrossRef] [PubMed]
88. Driessen, C.; Hirv, K.; Rink, L.; Kirchner, H. Induction of cytokines by zinc ions in human peripheral blood
mononuclear cells and separated monocytes. Lymphokine Cytokine Res. 1994, 13, 15–20. [PubMed]
89. Foote, J.W.; Delves, H.T. Albumin bound and alpha 2-macroglobulin bound zinc concentrations in the sera
of healthy adults. J. Clin. Pathol. 1984, 37, 1050–1054. [CrossRef] [PubMed]
90. Castell, J.V.; Gómez-Lechón, M.J.; David, M.; Andus, T.; Geiger, T.; Trullenque, R.; Fabra, R.; Heinrich, P.C.
Interleukin-6 is the major regulator of acute phase protein synthesis in adult human hepatocytes. FEBS Lett.
1989, 242, 237–239. [CrossRef]
91. Jang, J.Y.; Shim, H.; Lee, S.H.; Lee, J.G. Serum selenium and zinc levels in critically ill surgical patients.
J. Crit. Care 2014, 29, 317.e5–317.e8. [CrossRef] [PubMed]
92. Trame, S.; Wessels, I.; Haase, H.; Rink, L. A short 18 items food frequency questionnaire biochemically
validated to estimate zinc status in humans. J. Trace Elem. Med. Biol. 2018, 49, 285–295. [CrossRef] [PubMed]
93. Sukhavasi, S.; Jothimuthu, P.; Papautsky, I.; Beyette, F.R. Development of a Point-of-Care device to quantify
serum zinc to aid the diagnosis and follow-up of pediatric septic shock. In Proceedings of the 2011 Annual
International Conference of the IEEE Engineering in Medicine and Biology Society, Boston, MA, USA,
30 August–3 September 2011; pp. 3676–3679. [CrossRef]
94. Zerhusen, B.; de Silva, G.; Pei, X.; Papautsky, I.; Beyette, F.R. Rapid quantification system for zinc in blood
serum. In Proceedings of the 2013 IEEE 56th International Midwest Symposium on Circuits and Systems
(MWSCAS), Columbus, OH, USA, 4–7 August 2013; pp. 400–403.
95. Maret, W. Analyzing free zinc(II) ion concentrations in cell biology with fluorescent chelating molecules.
Metallomics 2015, 7, 202–211. [CrossRef] [PubMed]
96. Bornhorst, J.; Kipp, A.P.; Haase, H.; Meyer, S.; Schwerdtle, T. The crux of inept biomarkers for risks and
benefits of trace elements. TrAC Trends Anal. Chem. 2017. [CrossRef]
© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
262
nutrients
Review
Immunomodulatory Protein Hydrolysates and
Their Application
Mensiena B. G. Kiewiet *, Marijke M. Faas and Paul de Vos
Immunoendocrinology, Division of Medical Biology, Department of Pathology and Medical Biology,
University Medical Center Groningen, University of Groningen, Hanzeplein 1, 9700 RB Groningen,
The Netherlands; m.m.faas@umcg.nl (M.M.F.); p.de.vos@umcg.nl (P.d.V.)
* Correspondence: m.b.g.kiewiet@umcg.nl; Tel.: +31-50-361-0109
1. Introduction
Protein hydrolysates are commonly used as an alternative protein source in commercial products.
They consist of a mixture of different proteins and peptides which is formed by the hydrolysis
of intact proteins. During this process, peptide bonds of intact proteins are broken (Figure 1A),
which results in the formation of a range of peptides of different sizes. Depending on their properties,
protein hydrolysates are applied in different products. Mildly hydrolyzed proteins are, for example,
used in clinical and sport nutrition to support digestibility, while extensively hydrolyzed proteins are
used in infant formulas as a hypo-allergenic alternative for intact cow’s milk proteins (Figure 1B).
Furthermore, protein hydrolysates are recognized as a potent source of bioactive peptides.
Different peptides with, for example, anti-thrombotic, anti-hypertensive, anti-microbial, anti-cancer,
anti-oxidative, and many immunomodulatory effects have been identified [1]. Consuming protein
hydrolysates containing these peptides might be helpful in the management of many western
diseases [2,3]. Since many of these diseases are immune-related, immunomodulatory products have
gained special attention from both academical and industrial researchers for the management and
amelioration of, for example, inflammatory bowel diseases, allergies, and diabetes [4,5].
Figure 1. The process of protein hydrolysis and its products. (A) chemical reaction of protein hydrolysis;
(B) different hydrolysates serve different purposes.
264
Nutrients 2018, 10, 904
protein hydrolysates from the same protein source also have remarkably different immunomodulating
properties. Phagocytosis modulation was source dependent as well. Most protein hydrolysates were
found to increase macrophage phagocytosis capacity in vitro, for example, soy, egg, wheat, and casein
hydrolysates had such an effect [9,12,16,17]. However, a rice protein hydrolysate was found to inhibit
the phagocytic activity of RAW264.7 macrophages [18].
These early in vitro experiments investigating the immune effects of protein hydrolysates have
led to many more in vitro and in vivo studies on the effects of protein hydrolysates on immunity.
In the sections below, these in vivo studies are reviewed (an overview of the studies is given in Table 1).
In this part, a distinction was made between local intestinal immune effects and systemic effects.
All effects discussed below are also visualized in Figure 2. To start with, protein hydrolysates have an
effect on the epithelial cells aligning the gut and by that induce crosstalk between the epithelial cells
and immune cells [5].
Figure 2. Overview of immune effects induced by protein hydrolysates on the (A) intestinal epithelial
cells; (B) intestinal immune cells; (C) mesenteric lymph nodes; (D) systemic immune system.
265
Table 1. Overview of hydrolysates and their immune effects.
266
Increase of the IL-6 secretion by small intestinal epithelial
Egg yolk digests Pepsin cells, increase in IgA+ cells, orchestrating the Th1/Th2 mouse [24]
response.
Increase of secretory immunoglobulin A in the gut.
Pepsin hydrolysate increased the splenic NK cell
cytotoxicity, macrophage phagocytosis and level of serum
Common carp egg hydrolysate Pepsin, alcalase immunoglobulin A (IgA). S-IgA in the gut was significantly mouse [25]
enhanced by pepsin and alcalase hydrolysates. Trypsin
hydrolysate increased the percentages of CD4+ and CD8+
cells in the spleen.
Increased number of IgA+ cells in the small intestine lamina
Yellow field pea hydrolysate Thermolysin propria, accompanied by an increase in the number of IL-4+ , mouse [26]
IL-10+ , and IFNγ + cells.
Enhanced phagocytic activity of peritoneal macrophages,
increased number of IgA+ cells, and increased IL-4, IL-6,
Fermented pacific whiting protein Yeast mouse [27]
IL-10, IFNγ, and TNFα levels in the small intestine lamina
propria
Shark protein hydrolysate PeptibalTM Trypsin and Increase of small intestinal immunoglobulin A-producing
mouse [28]
(innoVactiv Inc) chymotrypsin cells and intestinal IL-6, TNFα, TGFβ, and IL-10
Table 1. Cont.
267
in MLN.
Decrease of inflammatory injury, as assessed by lower
extension of necrosis and damage score, myeloperoxidase, Rat (TNBS induced
bovine glycomacropeptide NA [33]
alkaline phosphatase, inducible nitric oxide synthase, IL- 1β, colitis)
TNFα, and IL-17.
Attenuated DSS-induced clinical symptoms, including
weight loss, mucosal and submucosal inflammation, crypt
distortion, and colon muscle thickening, and decreased
Pig (DSS induced
Egg white hydrolysate Aminopeptidase intestinal permeability and increased mucin gene [34]
colitis)
expression, reduced intestinal expression of
pro-inflammatory cytokines TNFα, IL-6, IL-1β, IFNγ, IL-8,
and IL-17.
Increased number of IL-12+ CD11b+ in spleens, increased
Rhizopus oryzae neutral
Soybean protein hydrolysate cytotoxic activity of spleen cells, increased Igh-4, Aqp8, Mouse [35]
protease preparation
Dmbt1, Slpi, and Mx1 in Peyer’s patch cells.
Increased Breg and Treg in the spleen, increased IgA+ B-cells
Partially hydrolyzed whey protein NA in the MLN, increased Th1, activated Treg and activated Mouse [36]
Th17 cells in the Peyer’s patches
Table 1. Cont.
the MLN.
Reduced acute allergic skin response and mast cell
Partial whey hydrolysate NA degranulation after whey challenge, increased Foxp3+ Mouse [38]
regulatory T-cell numbers in the MLN.
oyster peptide-based enteral Bromelain, pepsin, Enhanced spleen lymphocyte proliferation and of NK
Mouse [39]
nutrition formula trypsin cell activity
Casein hydrolysate Trypsin Phagocytosing capacity of phagocytic cells was increased Mouse [17]
Improved the level of hemolysin in serum, and enhanced
phagocytosis of macrophages. In ovalbumin-sensitized
mice, the milk protein hydrolysates reduced IgE levels,
Milk protein hydrolysate ICR mouse [40]
reduced IL-4 in serum, reduced the release of histamine and
bicarbonate in peritoneal mast cells, and enhanced TGFβ
levels.
268
Enhanced lymphocyte proliferation capacity increased
number of plaque-forming cells, increased NK cell activity,
Chum salmon oligopeptide preparation Complex protease increased percentage of CD4+ T helper (Th) cells in spleen ICR mouse [41]
and secretion of Th1 (IL-2, IFNγ) and Th2 (IL-5, IL-6) type
cell cytokines.
Increased weight of the spleen and thymus and enhanced
Tuna cooking drip hydrolysate Enzyme A and B the proliferation of splenocytes. Increased production of Mouse [42]
IL-10 and IL-2. Increased serum IgG1 and IgG2a levels.
Soy protein hydrolysate Pepsin Increased serum IgA and IgG levels Rat [7]
Theroase, bioprase, Total lymphocyte and granulocyte numbers were altered,
Soy protein hydrolysate Human [43]
Sumizyme FP and the numbers of CD11b+ cells and CD56+ cells increased.
Wheat gluten hydrolysate NA NK cell activity increased significantly Human [44]
Fish protein hydrolysate (Amizate) NA No effects observed Human [45]
Nutrients 2018, 10, 904
269
Nutrients 2018, 10, 904
Anti-inflammatory properties are another often observed feature of protein hydrolysates. This is a
characteristic mainly attributed to the hydrolysates of bovine milk. To study these effects, experimental
models with colitis and ileitis have been used. In all cases, oral pretreatment of the animals with casein
hydrolysate or a casein glycomacropeptide reduced damage to the intestine, leading to a decrease in
weight loss [30–33]. They also observed a decrease in the pro-inflammatory cytokines IL-1β, IL-17,
TNFα, and IFNγ, while in one study, an increase of the regulatory cytokine IL-10 was observed [31].
Similar outcomes were observed when egg white peptides were administered to pigs with dextran
sodium sulphate (DSS) induced colitis [34].
The effects of protein hydrolysates in the Peyer’s patches (PP), which are part of the gut-associated
lymphoid tissue, have been studied to a lesser extent but are essential as this tissue is the central
immune sampling and signaling site in the gut. Egusa et al. studied the effects of a specific soybean
protein digest in the PP [35]. After five weeks of feeding mice a soybean hydrolysate enriched diet,
the genome wide gene expression of PP derived cells was examined. They found that several genes
related to innate immunity and host defense were upregulated. They observed the upregulation of
Igh-4 and Aqp8, which enhance phagocytosis, and of Dmbt1, Slpi, and Mx1, which are anti-bacterial
and anti-viral components [52–54]. When looking at adaptive responses in the PP, we found in our
own study that sensitization with a partially hydrolyzed whey protein prevented the increase of Th1,
Th17, and regulatory T cells (Treg) in the PP after a challenge with intact whey [36].
270
Nutrients 2018, 10, 904
Similar to results after direct in vitro stimulation of splenocytes, oral administration of protein
hydrolysate derived from oyster, tuna cooking drip, salmon, and multiple common carp egg
hydrolysates increased ex vivo splenocyte proliferation [25,39,41,42]. Some of these studies looked
into the cell types present in the spleens of these animals in more detail. Alcalase common carp egg
hydrolysate was found to increase CD4+ and CD8+ cells in the spleen [25], while salmon hydrolysate
only increased CD4+ [41]. These cells were expected to be both Th1 and Th2 cells, since both Th1 (IL-2,
IFNγ) and Th2 (IL-4, IL-5) cytokines were detected in the blood of these animals. These effects might
differ between protein hydrolysates, since a tuna cooking drip hydrolysate was found to increase IL-2
and IL-10 [42], and sensitization with a partial whey hydrolysate increased Treg in the spleen [36].
Together with an increase in Treg, an increase in regulatory B cells (Breg) was observed in the
spleen after whey hydrolysate administration [36]. Breg are increasingly recognized to be important in
regulatory immune responses and induce the differentiation of T cells into Treg [59]. Another study
also found an increase in IL-10 producing Breg when inducing oral tolerance using intact casein in
casein-allergic mice [60]. Here, casein might be digested in the intestine of the mice, after which the
newly formed bioactive peptide(s) derived from casein increased Breg. An adoptive transfer of these
Breg could even prevent the onset of allergy in recipients [60], demonstrating the importance of this
cell type in tolerance induction.
Other evidence that hydrolysate consumption affects B cell responses is the observation of
increased antibody levels observed in the blood. A soy protein hydrolysate was found to induce IgG
and IgA in the blood of rats [7], while the common carp egg hydrolysate increased IgA in mice [25].
Therefore, it is likely that protein hydrolysates not only affect B cell differentiation, but also induce
class-switching and antibody production.
Only a few studies investigated the effects of protein hydrolysates on immune parameters in the
blood of humans. The effect of a single dose of soybean hydrolysate was studied in a small group of
volunteers, and it was found to change leukocyte numbers and increase granulocytes. More specifically,
it significantly increased CD11b+ (macrophages and/or dendritic cells) and CD56+ cells (NK cells)
in blood [43]. When nine subjects were fed a wheat hydrolysate for six days, an increase in NK
cell activity was observed [44]. A larger group of undernourished Indian children was given a fish
hydrolysate for 120 days. After this period, the CD4/CD8 ratio and antibody levels were measured,
but no significant differences were detected [45]. These studies show that protein hydrolysates
have immunomodulatory effects in humans, which are similar to some of the previously mentioned
in vitro and in vivo effects, although the protein hydrolysates used are different. However, in these
studies, only a few immunological parameters were measured, and/or only included a small group of
volunteers, which makes it difficult to draw firm conclusions. Well-designed, extensive human studies
are lacking at the moment, but are needed to better understand the effects of protein hydrolysates
in humans.
271
Nutrients 2018, 10, 904
peptides in the hydrolysate based on their physicochemical properties, mainly size, hydrophobicity,
or a combination thereof. Techniques which are often used to characterize peptide composition are,
for example sodium, dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE), size exclusion
chromatography, and High Performance Liquid Chromatography (HPLC) [63]. All these techniques
are used to obtain a molecular weight distribution, which gives an overall indication of the hydrolysate
composition based on size. One can focus on a specific fraction of the protein hydrolysates by choosing
the most appropriate technique and optimizing the settings used.
HPLC, especially Reverse Phase-HPLC, has, for a long time, been found to be useful in separating
peptides from protein hydrolysates based on their size and hydrophobicity [64]. When this method
is coupled to a mass spectrometer (MS), it is also possible to determine the amino acid sequence of
the detected peptides [65]. In this way, a very detailed characterization of the hydrolysate can be
obtained, which can be used to identify structure-effector relationships between, for example, immune
effects and specific peptides in a hydrolysate. However, protein hydrolysates consist of thousands of
different peptides, which will all be detected by HPLC-MS when a complete hydrolysate is analyzed.
Therefore, studies aiming for the identification of bioactive peptides often compare the bioactivity
of size-based fractions of the hydrolysate first [66,67], after which the bioactive fraction alone can be
further analyzed using HPLC-MS. Ultimately, the peptides present in the bioactive fraction should be
tested individually for bioactivity in order to obtain the peptide responsible for the observed effect [65].
Using this and similar methods has led to the discovery of a range of bioactive peptides from
protein mixtures, as reviewed by Sanchez et al., and Lafarga et al., [1,68]. However, most of the identified
bioactive peptides possess anti-hypertensive, anti-microbial, or anti-oxidative effects, but peptides with
immunomodulatory properties have not widely been identified yet. Comparing the characteristics of the
immunomodulatory peptides that have been identified generated new knowledge about which peptide
properties are associated with immune effects. It is known that immunomodulatory peptides are mostly
two to 20 amino acids long and hydrophobic [61]. Chalamaiah et al., concluded by listing known
peptides with immunomodulatory effects that glycine (Gly), valine (Val), leucine (Leu), proline (Pro),
phenylalanine (Phe), negatively charged amino acid glutamic acid (Glu), and aromatic amino acid
tyrosine (Tyr) were most frequently present in peptides with immune effects [69].
Recently, we also found that larger fractions in whey and soy protein hydrolysates can have
immunomodulatory effects. These fractions have a size of over 1000 kDa and were composed of
aggregates which are formed during the hydrolysis process [66]. In this process, the proteins are
heated, which is a known cause of protein denaturation and aggregation [70]. By performing PAGE
under different conditions, it was found that these aggregates were formed due to electrostatic forces
and disulfide bridges between single proteins and induce responses in human dendritic cells. The fact
that these aggregates were found in both a whey and soy hydrolysate suggests that aggregate formation
is not protein source specific and might be present in a wide range of different protein hydrolysates.
272
Nutrients 2018, 10, 904
Figure 3. Overview of mechanisms described in the literature via which peptides can exert
immunomodulatory effects in the cell. Peptides can (A) directly stimulate receptors; (B) be taken
up in the cell via a peptide transporter and interfere with inflammatory signaling pathways; or (C) be
taken up into the cell via endocytosis and inhibit inflammatory signaling pathways.
9. Receptor Binding
One of the most studied receptor types in immune signaling is Toll-like receptors (TLRs). This is a
family of pathogen recognition receptors. They are not only expressed by most immune cells [71,72],
but also by epithelial cells [73]. Multiple protein hydrolysates were found to affect TLRs, but the
effects were very hydrolysate dependent. By studying a range of cow’s milk hydrolysates in a TLR
reporter cell platform, we previously found that especially mildly hydrolyzed whey hydrolysates were
able to activate multiple TLRs, including TLR2, 3, 4, 5, 7, 8, and 9 [74]. This activation does lead to
the production of TNFα, IL-10, and IL-8 in human peripheral blood mononuclear cells (PBMCs).
The protein source of the hydrolysates was crucial for its final effects, as it was observed that
casein hydrolysates only inhibited TLR activation. Which TLRs were inhibited also differed per
hydrolysate, but TLR5 and 9 were the most profoundly inhibited. Other studies focused mainly on
TLR2 and 4. Tobita et al., described that a casein phosphopeptide was able to induce proliferation
and IL-6 production in CD19+ cells from mice after stimulation in vitro. This effect was gone after the
administration of an anti-TLR4 antibody, suggesting that the effects were induced via TLR4 [75].
A primary culture of murine intestinal epithelial cells was also thought to secrete IL-6 via the
stimulation of both TLR2 and TLR4 when the mice were fed with yellow pea hydrolysate [26] and shark
protein hydrolysate [28]. A pressurized whey hydrolysate was able to suppress lipopolysaccharide
(LPS) induced IL-8 production in respiratory epithelial cell lines, likely via binding to TLR4 [76].
There is evidence that other transmembrane receptors are also involved in immune effects by
protein hydrolysates. Tsuruki et al., described that a peptide derived from the soybean β-conglycinin
A’ subunit, which stimulated phagocytosis in human neutrophils, showed a low affinity for the
N-formyl-methionyl-leucyl-phenylalanine (fMLP) receptor [77]. The phagocytosis stimulating effect
273
Nutrients 2018, 10, 904
disappeared when this receptor was blocked. The authors discuss that its low affinity for the fMLP
receptor allows it to stimulate the immune response in a safe way, without causing inflammation.
Peptides derived from many different food protein sources are known to bind opioid receptors [78].
Although endogenous opioid peptides have a main function as neurotransmitters, they are also known to
modulate innate and acquired immune responses [79]. Opioid receptor signaling can, for example, skew T
cell differentiation, increase antibody production in B cells, and affect phagocytosis in macrophages [80–82].
These effects have also been described in immune cells after hydrolysate administration. Therefore,
it cannot be excluded that protein hydrolysates also modulate the immune system via opioid receptors.
11. Endocytosis
Once immunomodulatory peptides are taken up in epithelial or immune cells, they can exert their
effects by interfering with signaling pathways. However, not all peptides and protein can be internalized
via the PepT1 transporter, since this transporter is specific for di- and tripeptides. Larger food derived
peptides, which are too large for the PepT1 transporter, can also be taken up into the cell by fluid phase
endocytosis (Figure 3C), which is a non-specific form of vesicle mediated internalization. In this type of
endocytosis, hydrophobic interactions between the peptide and the cell membrane are involved in the
internalization of the peptide [89]. Differences in physicochemical properties of peptides, including size,
hydrophobicity, and charge determine the kinetics of uptake of individual peptides.
The involvement of this type of peptide uptake in immunomodulation by peptides was confirmed
by Regazzo et al., [90]. They showed that a relatively large, hydrophobic casein peptide which
showed multiple stimulating effects in immune cells, could be taken up in a layer of Caco2 cells via
endocytosis. They did not see a difference in casein peptide uptake when they used an inhibitor for
274
Nutrients 2018, 10, 904
the PepT1 transporter or cytochalasin D to open the tight junctions (and increase the paracellular
route), but found a significant inhibitory effect on peptide uptake after treatment with wortmannin,
which inhibits endocytosis [90]. When the peptide is translocated over the epithelial cells via this
mechanism, it may affect immune cell functioning in the lamina propria or in the blood.
A study investigating the well-characterized soy peptide lunasin showed that endocytosis is
also used by immune cells to take up immunomodulatory peptides [91]. Lunasin is a 43-amino
acid peptide which was shown to interact with the αVβ3 integrin, which led to inhibiting αVβ3
integrin-mediated pro-inflammatory markers and to downregulation of the Akt-mediated NF-κB
pathway. By using different inhibitors, it was found that lunasin was mainly taken up by endocytic
mechanisms that involve integrin signaling, clathrin-coated structures, and macropinosomes [91].
Interestingly, this lunasin uptake was increased under inflammatory conditions.
275
Nutrients 2018, 10, 904
Figure 4. Summary of the application possibilities of protein hydrolysates. These hydrolysates are
currently being used in sport nutrition, clinical nutrition, and infant formula, mainly because of their
good digestibility and hypoallergenicity. Recent research indicates that specific protein hydrolysates
could optimize the current products in multiple ways. Also, there is evidence that new protein
hydrolysate products could be beneficial for specific target groups.
Figure 5. Overview of T helper cell subsets and their interactions and their relation to specific diseases.
276
Nutrients 2018, 10, 904
277
Nutrients 2018, 10, 904
due to stress are increased pro-inflammatory cytokines, more Th2-related cytokines, and changes in
leukocyte number and distribution [123–126]. As described above, specific protein hydrolysates are able
to counteract these effects. Therefore, functional food containing protein hydrolysates may contribute to
healthy immunity in people experiencing significant stress levels.
15. Conclusions
A wide range of protein hydrolysates have immunomodulatory capacities. However, before
protein hydrolysates can serve as functional foods, physicochemical approaches to identify the protein
sequence(s) are needed to be able to design effective protein hydrolysates. Also, specific target groups
have to be identified. In this way, specific protein hydrolysates can be designed to ameliorate, delay,
or prevent the onset of a wide variety of Western immune-related conditions.
Author Contributions: Writing-Original Draft Preparation, M.B.G.K.; Writing-Review & Editing, M.M.F.
and P.d.V.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Sánchez, A.; Vázquez, A. Bioactive peptides: A review. Food Qual. Saf. 2017, 1, 29–46. [CrossRef]
2. Li-Chan, E.C.Y. Bioactive peptides and protein hydrolysates: Research trends and challenges for application
as nutraceuticals and functional food ingredients. Curr. Opin. Food Sci. 2015, 1, 28–37. [CrossRef]
3. Dhaval, A.; Yadav, N.; Purwar, S. Potential Applications of Food Derived Bioactive Peptides in Management
of Health. Int. J. Pept. Res. Ther. 2016, 22, 377–398. [CrossRef]
4. Bouglé, D.; Bouhallab, S. Dietary bioactive peptides: Human studies. Crit. Rev. Food Sci. Nutr. 2015,
57, 335–343. [CrossRef] [PubMed]
5. Kiewiet, M.B.G.; Gros, M.; van Neerven, R.J.J.; Faas, M.M.; de Vos, P. Immunomodulating properties of
protein hydrolysates for application in cow’s milk allergy. Pediatr. Allergy Immunol. 2015, 26, 206–217.
[CrossRef] [PubMed]
6. Parker, F.; Migliore-Samour, D.; Floch, F.; Zerial, A.; Werner, G.H.; Jollès, J.; Casaretto, M.; Zahn, H.; Jollès, P.
Immunostimulating hexapeptide from human casein: Amino acid sequence, synthesis and biological
properties. Eur. J. Biochem. 1984, 145, 677–682. [CrossRef] [PubMed]
7. Ashaolu, T.J.; Yanyiam, N.; Yupanqui, C.T. Immunomodulatory effects of pepsin-educed soy protein
hydrolysate in rats and murine cells. Funct. Foods Health Dis. 2017, 7, 889–900.
8. Caron, S.; Samanez, C.H.; Dehondt, H.; Ploton, M.; Briand, O.; Lien, F.; Dorchies, E.; Dumont, J.; Postic, C.;
Cariou, B.; et al. Farnesoid X receptor inhibits the transcriptional activity of carbohydrate response element
binding protein in human hepatocytes. Mol. Cell. Biol. 2013, 33, 2202–2211. [CrossRef] [PubMed]
9. Kong, X.; Guo, M.; Hua, Y.; Cao, D.; Zhang, C. Enzymatic preparation of immunomodulating hydrolysates
from soy proteins. Bioresour. Technol. 2008, 99, 8873–8879. [CrossRef] [PubMed]
10. Mao, X.Y.; Yang, H.Y.; Song, J.P.; Li, Y.H.; Ren, F.Z. Effect of yak milk casein hydrolysate on Th1/Th2
cytokines production by murine spleen lymphocytes in vitro. J. Agric. Food Chem. 2007, 55, 638–642.
[CrossRef] [PubMed]
11. Rodríguez-Carrio, J.; Fernández, A.; Riera, F.A.; Suárez, A. Immunomodulatory activities of whey
β-lactoglobulin tryptic-digested fractions. Int. Dairy J. 2014, 34, 65–73. [CrossRef]
12. Wu, W.; Zhang, M.; Sun, C.; Brennan, M.; Li, H.; Wang, G.; Lai, F.; Wu, H. Enzymatic preparation of
immunomodulatory hydrolysates from defatted wheat germ (Triticum Vulgare) globulin. Int. J. Food
Sci. Technol. 2016, 51, 2556–2566. [CrossRef]
13. He, X.Q.; Cao, W.H.; Pan, G.K.; Yang, L.; Zhang, C.H. Enzymatic hydrolysis optimization of Paphia undulata
and lymphocyte proliferation activity of the isolated peptide fractions. J. Sci. Food Agric. 2015, 95, 1544–1553.
[CrossRef] [PubMed]
14. Lozano-Ojalvo, D.; Molina, E.; López-Fandiño, R. Hydrolysates of egg white proteins modulate T- and B-cell
responses in mitogen-stimulated murine cells. Food Funct. 2016, 7, 1048–1056. [CrossRef] [PubMed]
278
Nutrients 2018, 10, 904
15. Sutas, Y.; Soppi, E.; Korhonen, H.; Syvaoja, E.L.; Saxelin, M.; Rokka, T.; Isolauri, E. Suppression of lymphocyte
proliferation in vitro by bovine caseins hydrolyzed with Lactobacillus casei GG-derived enzymes. J. Allergy
Clin. Immunol. 1996, 98, 216–224. [CrossRef]
16. Meram, C.; Wu, J. Anti-inflammatory effects of egg yolk livetins (α, β, and γ-livetin) fraction and its
enzymatic hydrolysates in lipopolysaccharide-induced RAW 264.7 macrophages. Food Res. Int. 2017,
100, 449–459. [CrossRef] [PubMed]
17. Kazlauskaite, J.; Biziulevicius, G.A.; Zukaite, V.; Biziuleviciene, G.; Miliukiene, V.; Siaurys, A. Oral tryptic
casein hydrolysate enhances phagocytosis by mouse peritoneal and blood phagocytic cells but fails to
prevent induced inflammation. Int. Immunopharmacol. 2005, 5, 1936–1944. [CrossRef] [PubMed]
18. Wen, L.; Chen, Y.; Zhang, L.; Yu, H.; Xu, Z.; You, H.; Cheng, Y. Rice protein hydrolysates (RPHs) inhibit
the LPS-stimulated inflammatory response and phagocytosis in RAW264.7 macrophages by regulating the
NF-κB signaling pathway. RSC Adv. 2016, 6, 71295–71304. [CrossRef]
19. Visser, J.; Lammers, K.; Hoogendijk, A.; Boer, M.; Brugman, S.; Beijer-Liefers, S.; Zandvoort, A.; Harmsen, H.;
Welling, G.; Stellaard, F.; et al. Restoration of impaired intestinal barrier function by the hydrolysed casein
diet contributes to the prevention of type 1 diabetes in the diabetes-prone BioBreeding rat. Diabetologia 2010,
53, 2621–2628. [CrossRef] [PubMed]
20. Visser, J.; Bos, N.; Harthoorn, L.; Stellaard, F.; Beijer-Liefers, S.; Rozing, J.; van Tol, E.A.F. Potential mechanisms
explaining why hydrolyzed casein-based diets outclass single amino acid-based diets in the prevention
of autoimmune diabetes in diabetes-prone BB rats. Diabete Metab. Res. Rev. 2012, 28, 505–513. [CrossRef]
[PubMed]
21. Plaisancié, P.; Claustre, J.; Estienne, M.; Henry, G.; Boutrou, R.; Paquet, A.; Léonil, J. A novel bioactive peptide
from yoghurts modulates expression of the gel-forming MUC2 mucin as well as population of goblet cells
and Paneth cells along the small intestine. J. Nutr. Biochem. 2013, 24, 213–221. [CrossRef] [PubMed]
22. Thoreux, K.; Balas, D.; Bouley, C.; Senegas-Balas, F. Diet Supplemented with Yoghurt or Milk Fermented by
Lactobacillus casei DN-114 001 Stimulates Growth and Brush-Border Enzyme Activities in Mouse Small
Intestine. Digestion 1998, 59, 349–359. [CrossRef] [PubMed]
23. Vinderola, G.; Matar, C.; Perdigón, G. Milk fermentation products of L. helveticus R389 activate calcineurin
as a signal to promote gut mucosal immunity. BMC Immunol. 2007, 8, 19. [CrossRef] [PubMed]
24. Nelson, R.; Katayama, S.; Mine, Y.; Duarte, J.; Matar, C. Immunomodulating effects of egg yolk low lipid
peptic digests in a murine model. Food Agric. Immunol. 2007, 18, 1–15. [CrossRef]
25. Chalamaiah, M.; Hemalatha, R.; Jyothirmayi, T.; Diwan, P.V.; Bhaskarachary, K.; Vajreswari, A.;
Ramesh Kumar, R.; Dinesh Kumar, B. Chemical composition and immunomodulatory effects of enzymatic
protein hydrolysates from common carp (Cyprinus carpio) egg. Nutrition 2015, 31, 388–398. [CrossRef]
[PubMed]
26. Ndiaye, F.; Vuong, T.; Duarte, J.; Aluko, R.E.; Matar, C. Anti-oxidant, anti-inflammatory and
immunomodulating properties of an enzymatic protein hydrolysate from yellow field pea seeds. Eur. J. Nutr.
2012, 51, 29–37. [CrossRef] [PubMed]
27. Duarte, J.; Vinderola, G.; Ritz, B.; Perdigón, G.; Matar, C. Immunomodulating capacity of commercial fish
protein hydrolysate for diet supplementation. Immunobiology 2006, 211, 341–350. [CrossRef] [PubMed]
28. Mallet, J.F.; Duarte, J.; Vinderola, G.; Anguenot, R.; Beaulieu, M.; Matar, C. The immunopotentiating effects
of shark-derived protein hydrolysate. Nutrition 2014, 30, 706–712. [CrossRef] [PubMed]
29. LeBlanc, J.; Fliss, I.; Matar, C. Induction of a Humoral Immune Response following an Escherichia
coli O157:H7 Infection with an Immunomodulatory Peptidic Fraction Derived from Lactobacillus
helveticus-Fermented Milk. Clin. Diagn. Lab. Immunol. 2004, 11, 1171–1181. [CrossRef] [PubMed]
30. Daddaoua, A.; Puerta, V.; Zarzuelo, A.; Suárez, M.D.; Sánchez de Medina, F.; Martínez-Augustin, O.
Bovine Glycomacropeptide Is Anti-Inflammatory in Rats with Hapten-Induced Colitis. J. Nutr. 2005,
135, 1164–1170. [CrossRef] [PubMed]
31. Espeche Turbay, M.B.; De Leblanc, A.D.M.; Perdigón, G.; Savoy de Giori, G.; Hebert, E.M.
β-Casein hydrolysate generated by the cell envelope-associated proteinase of Lactobacillus delbrueckii
ssp. lactis CRL 581 protects against trinitrobenzene sulfonic acid-induced colitis in mice. J. Dairy Sci. 2012,
95, 1108–1118. [CrossRef] [PubMed]
279
Nutrients 2018, 10, 904
32. Ortega-Gonzalez, M.; Capitan-Canadas, F.; Requena, P.; Ocon, B.; Romero-Calvo, I.; Aranda, C.; Suarez, M.D.;
Zarzuelo, A.; de Medina, F.S.; Martinez-Augustin, O. Validation of bovine glycomacropeptide as an intestinal
anti-inflammatory nutraceutical in the lymphocyte-transfer model of colitis. Br. J. Nutr. 2014, 111, 1202–1212.
[CrossRef] [PubMed]
33. Requena, P.; Daddaoua, A.; Martínez-Plata, E.; González, M.; Zarzuelo, A.; Suárez, M.D.; Sánchez de Medina, F.;
Martínez-Augustin, O. Bovine glycomacropeptide ameliorates experimental rat ileitis by mechanisms involving
downregulation of interleukin 17. Br. J. Pharmacol. 2009, 154, 825–832. [CrossRef] [PubMed]
34. Lee, M.; Kovacs-Nolan, J.; Archbold, T.; Fan, M.Z.; Juneja, L.R.; Okubo, T.; Mine, Y. Therapeutic potential
of hen egg white peptides for the treatment of intestinal inflammation. J. Funct. Foods 2009, 1, 161–169.
[CrossRef]
35. Egusa, S.; Otani, H. Soybean protein fraction digested with neutral protease preparation, “Peptidase R”,
produced by Rhizopus oryzae, stimulates innate cellular immune system in mouse. Int. Immunopharmacol.
2009, 9, 931–936. [CrossRef] [PubMed]
36. Kiewiet, M.; van Esch, B.; Garssen, J.; Faas, M.; de Vos, P. Partially hydrolyzed whey proteins prevent clinical
symptoms in a cow’s milk allergy mouse model and enhance regulatory T and B cell frequencies. Mol. Nutr.
Food Res. 2017, 61. [CrossRef] [PubMed]
37. Meulenbroek, L.A.P.M.; van Esch, B.C.A.M.; Hofman, G.A.; den Hartog Jager, C.F.; Nauta, A.J.;
Willemsen, L.E.M.; Bruijnzeel-Koomen, C.A.F.M.; Garssen, J.; van Hoffen, E.; Knippels, L.M.J. Oral treatment
with beta-lactoglobulin peptides prevents clinical symptoms in a mouse model for cow’s milk allergy.
Pediatr. Allergy Immunol. 2013, 24, 656–664. [CrossRef] [PubMed]
38. Van Esch, B.C.A.M.; Schouten, B.; de Kivit, S.; Hofman, G.A.; Knippels, L.M.J.; Willemsen, L.E.M.; Garssen, J.
Oral tolerance induction by partially hydrolyzed whey protein in mice is associated with enhanced numbers
of Foxp3(+) regulatory T-cells in the mesenteric lymph nodes. Pediatr. Allergy Immunol. 2011, 22, 820–826.
[CrossRef] [PubMed]
39. Cai, B.; Pan, J.; Wu, Y.; Wan, P.; Sun, H. Immune functional impacts of oyster peptide-based enteral nutrition
formula (OPENF) on mice: A pilot study. Chin. J. Oceanol. Limnol. 2013, 31, 813–820. [CrossRef]
40. Pan, D.D.; Wu, Z.; Liu, J.; Cao, X.Y.; Zeng, X.Q. Immunomodulatory and hypoallergenic properties of milk
protein hydrolysates in ICR mice. J. Dairy Sci. 2013, 96, 4958–4964. [CrossRef] [PubMed]
41. Yang, R.; Zhang, Z.; Pei, X.; Han, X.; Wang, J.; Wang, L.; Long, Z.; Shen, X.; Li, Y. Immunomodulatory effects
of marine oligopeptide preparation from Chum Salmon (Oncorhynchus keta) in mice. Food Chem. 2009,
113, 464–470. [CrossRef]
42. Kim, M.J.; Kim, K.B.W.R.; Sung, N.Y.; Byun, E.H.; Nam, H.S.; Ahn, D.H. Immune-enhancement effects of
tuna cooking drip and its enzymatic hydrolysate in Balb/c mice. Food Sci. Biotechnol. 2018, 27, 131–137.
[CrossRef]
43. Yimit, D.; Hoxur, P.; Amat, N.; Uchikawa, K.; Yamaguchi, N. Effects of soybean peptide on immune function,
brain function, and neurochemistry in healthy volunteers. Nutrition 2012, 28, 154–159. [CrossRef] [PubMed]
44. Horiguchi, N.; Horiguchi, H.; Suzuki, Y. Effect of wheat gluten hydrolysate on the immune system in healthy
human subjects. Biosci. Biotechnol. Biochem. 2005, 69, 2445–2449. [CrossRef] [PubMed]
45. Nesse, K.O.; Nagalakshmi, A.P.; Marimuthu, P.; Singh, M. Efficacy of a Fish Protein Hydrolysate in
Malnourished Children. Indian J. Clin. Biochem. 2011, 26, 360–365. [CrossRef] [PubMed]
46. Farhadi, A.; Banan, A.; Fields, J.; Keshavarzian, A. Intestinal barrier: An interface between health and disease.
J. Gastroenterol. Hepatol. 2003, 18, 479–497. [CrossRef] [PubMed]
47. Peterson, L.W.; Artis, D. Intestinal epithelial cells: Regulators of barrier function and immune homeostasis.
Nat. Rev. Immunol. 2014, 14, 141–153. [CrossRef] [PubMed]
48. Liu, F.; Poursine-Laurent, J.; Wu, H.; Link, D.C. Interleukin-6 and the granulocyte colony-stimulating
factor receptor are major independent regulators of granulopoiesis in vivo but are not required for lineage
commitment or terminal differentiation. Blood 1997, 90, 2583–2590. [PubMed]
49. Smith, K.A.; Maizels, R.M. IL-6 controls susceptibility to helminth infection by impeding Th2 responsiveness
and altering the Treg phenotype in vivo. Eur. J. Immunol. 2014, 44, 150–161. [CrossRef] [PubMed]
50. Hunter, C.A.; Jones, S.A. IL-6 as a keystone cytokine in health and disease. Nat. Immunol. 2015, 16, 448–457.
[CrossRef] [PubMed]
51. Pabst, O. New concepts in the generation and functions of IgA. Nat. Rev. Immunol. 2012, 12, 821–832.
[CrossRef] [PubMed]
280
Nutrients 2018, 10, 904
52. Ligtenberg, A.; Veerman, E.; Nieuw Amerongen, A.; Mollenhauer, J. Salivary agglutinin/glycoprotein-340/
DMBT1: A single molecule with variable composition and with different functions in infection, inflammation
and cancer. Biol. Chem. 2007, 388, 1275–1289. [CrossRef] [PubMed]
53. Fernie-King, B.; Seilly, D.; Binks, M.; Sriprakash, K.; Lachmann, P. Streptococcal DRS (distantly related to
SIC) and SIC inhibit antimicrobial peptides, components of mucosal innate immunity: A comparison of their
activities. Microbes Infect. 2007, 9, 300–307. [CrossRef] [PubMed]
54. Samuel, C.E. Antiviral actions of interferons. Clin. Microbiol. Rev. 2001, 14, 778–809. [CrossRef] [PubMed]
55. Macpherson, A.J.; Smith, K. Mesenteric lymph nodes at the center of immune anatomy. J. Exp. Med. 2006,
203, 497–500. [CrossRef] [PubMed]
56. Santiago, A.F.; Fernandes, R.M.; Santos, B.P.; Assis, F.A.; Oliveira, R.P.; Carvalho, C.R.; Faria, A.M.C. Role of
mesenteric lymph nodes and aging in secretory IgA production in mice. Cell. Immunol. 2008, 253, 5–10.
[CrossRef] [PubMed]
57. Chabance, B.; Marteau, P.; Rambaud, J.C.; Migliore-Samour, D.; Boynard, M.; Perrotin, P.; Guillet, R.; Jolles, P.;
Fiat, A.M. Casein peptide release and passage to the blood in humans during digestion of milk or yogurt.
Biochimie 1998, 80, 155–165. [CrossRef]
58. Dia, V.P.; Torres, S.; De Lumen, B.O.; Erdman, J.W.; Gonzalez De Mejia, E. Presence of lunasin in plasma of
men after soy protein consumption. J. Agric. Food Chem. 2009, 57, 1260–1266. [CrossRef] [PubMed]
59. Rosser, E.C.; Mauri, C. Regulatory B Cells: Origin, Phenotype, and Function. Immunity 2015, 42, 607–612.
[CrossRef] [PubMed]
60. Kim, A.R.; Kim, H.S.W.H.S.; Kim, D.K.; Nam, S.T.; Kim, H.S.W.H.S.; Park, Y.H.; Lee, D.; Lee, M.B.; Lee, J.H.;
Kim, B.; et al. Mesenteric IL-10-producing CD5(+) regulatory B cells suppress cow’s milk casein-induced
allergic responses in mice. Sci. Rep. 2016, 6, 19685. [CrossRef] [PubMed]
61. Agyei, D.; Ongkudon, C.M.; Wei, C.Y.; Chan, A.S.; Danquah, M.K. Bioprocess challenges to the isolation and
purification of bioactive peptides. Food Bioprod. Process. 2016, 98, 244–256. [CrossRef]
62. Rutherfurd, S.M. Methodology for determining degree of hydrolysis of proteins in Hydrolysates: A review.
J. AOAC Int. 2010, 93, 1515–1522. [PubMed]
63. Silvestre, M.P.C. Review of methods for the analysis of protein hydrolysates. Food Chem. 1997, 60, 263–271.
[CrossRef]
64. Lemieux, L.; Piot, J.-M.; Guillochon, D.; Amiot, J. Study of the efficiency of a mobile phase used
in size-exclusion HPLC for the separation of peptides from a casein hydrolysate according to their
hydrodynamic volume. Chromatographia 1991, 32, 499–504. [CrossRef]
65. Chen, H.-M.; Muramoto, K.; Yamauchi, F. Structural Analysis of Antioxidative Peptides from Soybean
.beta.-Conglycinin. J. Agric. Food Chem. 1995, 43, 574–578. [CrossRef]
66. Kiewiet, M.B.G.; Dekkers, R.; Ulfman, L.H.; Groeneveld, A.; de Vos, P.; Faas, M.M. Immunomodulating
protein aggregates in soy and whey hydrolysates and their resistance to digestion in an in vitro infant
gastrointestinal model: New insights in the mechanism of immunomodulatory hydrolysates. Food Funct.
2018, 9, 604–613. [CrossRef] [PubMed]
67. Mukhopadhya, A.; Noronha, N.; Bahar, B.; Ryan, M.T.; Murray, B.A.; Kelly, P.M.; O’Loughlin, I.B.;
O’Doherty, J.V.; Sweeney, T. The anti-inflammatory potential of a moderately hydrolysed casein and its 5 kDa
fraction in in vitro and ex vivo models of the gastrointestinal tract. Food Funct. 2015, 6, 612–621. [CrossRef]
[PubMed]
68. Lafarga, T.; Hayes, M. Bioactive protein hydrolysates in the functional food ingredient industry: Overcoming
current challenges. Food Rev. Int. 2017, 33, 217–246. [CrossRef]
69. Chalamaiah, M.; Yu, W.; Wu, J. Immunomodulatory and anticancer protein hydrolysates (peptides) from
food proteins: A review. Food Chem. 2018, 245, 205–222. [CrossRef] [PubMed]
70. Monahan, F.J.; German, J.B.; Kinsella, J.E. Effect of Ph and Temperature on Protein Unfolding and
Thiol-Disulfide Interchange Reactions during Heat-Induced Gelation of Whey Proteins. J. Agric. Food Chem.
1995, 43, 46–52. [CrossRef]
71. Gordon, S. Pattern recognition receptors: Doubling up for the innate immune response. Cell 2002,
111, 927–930. [CrossRef]
72. Michallet, M.-C.; Rota, G.; Maslowski, K.; Guarda, G. Innate receptors for adaptive immunity.
Curr. Opin. Microbiol. 2013, 16, 296–302. [CrossRef] [PubMed]
281
Nutrients 2018, 10, 904
73. Abreu, M.T. Toll-like receptor signalling in the intestinal epithelium: How bacterial recognition shapes
intestinal function. Nat. Rev. Immunol. 2010, 10, 131–143. [CrossRef] [PubMed]
74. Kiewiet, M.B.G.; Dekkers, R.; Gros, M.; van Neerven, R.J.J.; Groeneveld, A.; de Vos, P.; Faas, M.M.
Toll-like receptor mediated activation is possibly involved in immunoregulating properties of cow’s milk
hydrolysates. PLoS ONE 2017, 12, e0178191. [CrossRef] [PubMed]
75. Tobita, K.; Kawahara, T.; Otani, H. Bovine beta-casein (1-28), a casein phosphopeptide, enhances proliferation
and IL-6 expression of mouse CD19(+) cells via toll-like receptor 4. J. Agric. Food Chem. 2006, 54, 8013–8017.
[CrossRef] [PubMed]
76. Iskandar, M.M.; Lands, L.C.; Sabally, K.; Azadi, B.; Meehan, B.; Mawji, N.; Skinner, C.D.; Kubow, S.
High Hydrostatic Pressure Pretreatment of Whey Protein Isolates Improves Their Digestibility and
Antioxidant Capacity. Foods 2015, 4, 184–207. [CrossRef] [PubMed]
77. Tsuruki, T.; Kishi, K.; Takahashi, M.; Tanaka, M.; Matsukawa, T.; Yoshikawa, M. Soymetide,
an immunostimulating peptide derived from soybean beta-conglycinin, is an fMLP agonist. FEBS Lett. 2003,
540, 206–210. [CrossRef]
78. Stefanucci, A.; Mollica, A.; Macedonio, G.; Zengin, G.; Ahmed, A.A.; Novellino, E. Exogenous opioid
peptides derived from food proteins and their possible uses as dietary supplements: A critical review.
Food Rev. Int. 2018, 34, 70–86. [CrossRef]
79. Liang, X.; Liu, R.; Chen, C.; Ji, F.; Li, T. Opioid System Modulates the Immune Function: A Review.
Transl. Perioper. Pain Med. 2016, 1, 5–13. [CrossRef] [PubMed]
80. Börner, C.; Lanciotti, S.; Koch, T.; Höllt, V.; Kraus, J. μ opioid receptor agonist-selective regulation of
interleukin-4 in T lymphocytes. J. Neuroimmunol. 2013, 263, 35–42. [CrossRef] [PubMed]
81. Cheido, M.A.; Gevorgyan, M.M.; Zhukova, E.N. Comparative Evaluation of Opioid-Induced Changes in
Immune Reactivity of CBA Mice. Bull. Exp. Biol. Med. 2014, 156, 363–365. [CrossRef] [PubMed]
82. Tomassini, N.; Renaud, F.; Roy, S.; Loh, H.H. Morphine inhibits Fc-mediated phagocytosis through mu and
delta opioid receptors. J. Neuroimmunol. 2004, 147, 131–133. [CrossRef] [PubMed]
83. Dalmasso, G.; Charrier-Hisamuddin, L.; Thu Nguyen, H.T.; Yan, Y.; Sitaraman, S.; Merlin, D. PepT1-Mediated
Tripeptide KPV Uptake Reduces Intestinal Inflammation. Gastroenterology 2008, 134, 166–178. [CrossRef]
[PubMed]
84. Kovacs-Nolan, J.; Zhang, H.; Ibuki, M.; Nakamori, T.; Yoshiura, K.; Turner, P.V.; Matsui, T.; Mine, Y.
The PepT1-transportable soy tripeptide VPY reduces intestinal inflammation. Biochim. Biophys. Acta
Gen. Subj. 2012, 1820, 1753–1763. [CrossRef] [PubMed]
85. Oyama, M.; Van Hung, T.; Yoda, K.; He, F.; Suzuki, T. A novel whey tetrapeptide IPAV reduces interleukin-8
production induced by TNF-α in human intestinal Caco-2 cells. J. Funct. Foods 2017, 35, 376–383. [CrossRef]
86. Adibi, S.A. Regulation of expression of the intestinal oligopeptide transporter (Pept-1) in health and disease.
Am. J. Physiol. Liver Physiol. 2003, 285, G779–G788. [CrossRef] [PubMed]
87. Bermudez-Brito, M.; Sahasrabudhe, N.M.; Rösch, C.; Schols, H.A.; Faas, M.M.; de Vos, P. The impact of
dietary fibers on dendritic cell responses in vitro is dependent on the differential effects of the fibers on
intestinal epithelial cells. Mol. Nutr. Food Res. 2015, 59, 698–710. [CrossRef] [PubMed]
88. Charrier, L.; Driss, A.; Yan, Y.; Nduati, V.; Klapproth, J.-M.; Sitaraman, S.V.; Merlin, D. hPepT1 mediates
bacterial tripeptide fMLP uptake in human monocytes. Lab. Investig. 2006, 86, 490–503. [CrossRef] [PubMed]
89. Knipp, G.T.; Vander Velde, D.G.; Siahaan, T.J.; Borchardt, R.T. The effect of beta-turn structure on the passive
diffusion of peptides across Caco-2 cell monolayers. Pharm. Res. 1997, 14, 1332–1340. [CrossRef] [PubMed]
90. Regazzo, D.; Molle, D.; Gabai, G.; Tome, D.; Dupont, D.; Leonil, J.; Boutrou, R. The (193-209) 17-residues
peptide of bovine beta-casein is transported through Caco-2 monolayer. Mol. Nutr. Food Res. 2010,
54, 1428–1435. [CrossRef] [PubMed]
91. Cam, A.; Sivaguru, M.; Gonzalez de Mejia, E. Endocytic Mechanism of Internalization of Dietary Peptide
Lunasin into Macrophages in Inflammatory Condition Associated with Cardiovascular Disease. PLoS ONE
2013, 8, e72115. [CrossRef] [PubMed]
92. Kneepkens, C.M.F.; Meijer, Y. Clinical practice. Diagnosis and treatment of cow’s milk allergy. Eur. J. Pediatr.
2009, 168, 891–896. [CrossRef] [PubMed]
93. Turcanu, V.; Brough, H.A.; Du Toit, G.; Foong, R.-X.; Marrs, T.; Santos, A.F.; Lack, G. Immune mechanisms
of food allergy and its prevention by early intervention. Curr. Opin. Immunol. 2017, 48, 92–98. [CrossRef]
[PubMed]
282
Nutrients 2018, 10, 904
94. Berin, M.C.; Sampson, H.A. Review Mucosal Immunology of Food Allergy. Curr. Biol. 2013, 23, R389–R400.
[CrossRef] [PubMed]
95. Curotto de Lafaille, M.A.; Kutchukhidze, N.; Shen, S.; Ding, Y.; Yee, H.; Lafaille, J.J. Adaptive Foxp3+
Regulatory T Cell-Dependent and -Independent Control of Allergic Inflammation. Immunity 2008,
29, 114–126. [CrossRef] [PubMed]
96. Morgan, M.E.; Koelink, P.J.; Zheng, B.; den Brok, M.H.M.G.M.; van de Kant, H.J.; Verspaget, H.W.; Folkerts, G.;
Adema, G.J.; Kraneveld, A.D. Toll-like receptor 6 stimulation promotes T-helper 1 and 17 responses in
gastrointestinal-associated lymphoid tissue and modulates murine experimental colitis. Mucosal Immunol.
2014, 7, 1266–1277. [CrossRef] [PubMed]
97. Dolina, J.S.; Schoenberger, S.P. Toll-like receptor 9 is required for the maintenance of CD25+ FoxP3+ CD4+
Treg cells during Listeria monocytogenes infection. J. Immunol. 2017, 198 (Suppl. 1), 151.9.
98. Deutz, N.E.; Matheson, E.M.; Matarese, L.E.; Luo, M.; Baggs, G.E.; Nelson, J.L.; Hegazi, R.A.;
Tappenden, K.A.; Ziegler, T.R.; Grp, N.S. Readmission and mortality in malnourished, older, hospitalized
adults treated with a specialized oral nutritional supplement: A randomized clinical trial. Clin. Nutr. 2016,
35, 18–26. [CrossRef] [PubMed]
99. Elting, L.S.; Cooksley, C.; Chambers, M.; Cantor, S.B. The burdens of cancer therapy—Clinical and economic
outcomes of chemotherapy-induced mucositis. Cancer 2003, 98, 1531–1539. [CrossRef] [PubMed]
100. Kaczmarek, A.; Brinkman, B.M.; Heyndrickx, L.; Vandenabeele, P.; Krysko, D.V. Severity of
doxorubicin-induced small intestinal mucositis is regulated by the TLR-2 and TLR-9 pathways. J. Pathol.
2012, 226, 598–608. [CrossRef] [PubMed]
101. De Koning, B.A.E.; van Dieren, J.M.; Lindenbergh-Kortleve, D.J.; van der Sluis, M.; Matsumoto, T.;
Yamaguchi, K.; Einerhand, A.W.; Samsom, J.N.; Pieters, R.; Nieuwenhuis, E.E.S. Contributions of mucosal
immune cells to methotrexate-induced mucositis. Int. Immunol. 2006, 18, 941–949. [CrossRef] [PubMed]
102. Cario, E. Toll-like receptors in the pathogenesis of chemotherapy-induced gastrointestinal toxicity. Curr. Opin.
Support. Palliat. Care 2016, 10, 157–164. [CrossRef] [PubMed]
103. Villa, A.; Sonis, S.T. Mucositis: Pathobiology and management. Curr. Opin. Oncol. 2015, 27, 159–164.
[CrossRef] [PubMed]
104. Sahasrabudhe, N.M.; Beukema, M.; Tian, L.; Troost, B.; Scholte, J.; Bruininx, E.; Bruggeman, G.;
van den Berg, M.; Scheurink, A.; Schols, H.A.; et al. Dietary Fiber Pectin Directly Blocks Toll-Like Receptor
2–1 and Prevents Doxorubicin-Induced Ileitis. Front. Immunol. 2018, 9, 383. [CrossRef] [PubMed]
105. Crittenden, R.; Buckley, J.; Cameron-Smith, D.; Brown, A.; Thomas, K.; Davey, S.; Hobman, P. Functional
dairy protein supplements for elite athletes. Aust. J. Dairy Technol. 2009, 64, 133–138.
106. Foltz, M.; Ansems, P.; Schwarz, J.; Tasker, M.C.; Lourbakos, A.; Gerhardt, C.C. Protein hydrolysates induce
CCK release from enteroendocrine cells and act as partial agonists of the CCK1receptor. J. Agric. Food Chem.
2008, 56, 837–843. [CrossRef] [PubMed]
107. Kuehl, K.S.; Perrier, E.T.; Elliot, D.L.; Chesnutt, J.C. Efficacy of tart cherry juice in reducing muscle pain
during running: A randomized controlled trial. J. Int. Soc. Sports Nutr. 2010, 7, 1–6. [CrossRef] [PubMed]
108. Howatson, G.; McHugh, M.P.; Hill, J.A.; Brouner, J.; Jewell, A.P.; Van Someren, K.A.; Shave, R.E.;
Howatson, S.A. Influence of tart cherry juice on indices of recovery following marathon running. Scand. J.
Med. Sci. Sport. 2010, 20, 843–852. [CrossRef] [PubMed]
109. Koikawa, N.; Nakamura, A.; Ngaoka, I.; Aoki, K.; Sawaki, K.; Suzuki, Y. Delayed-onset muscle injury and its
modification by wheat gluten hydrolysate. Nutrition 2009, 25, 493–498. [CrossRef] [PubMed]
110. Hansen, M.; Bangsbo, J.; Bibby, B.M.; Madsen, K. Effect of whey protein hydrolysate on performance and
recovery of top-clas orienteering runners. Int. J. Sport Nutr. Exerc. Metab. 2014, 25, 97–109. [CrossRef]
[PubMed]
111. Buckley, J.D.; Thomson, R.L.; Coates, A.M.; Howe, P.R.C.; DeNichilo, M.O.; Rowney, M.K. Supplementation
with a whey protein hydrolysate enhances recovery of muscle force-generating capacity following eccentric
exercise. J. Sci. Med. Sport 2010, 13, 178–181. [CrossRef] [PubMed]
112. Cruzat, V.F.; Krause, M.; Newsholme, P. Amino acid supplementation and impact on immune function in
the context of exercise. J. Int. Soc. Sports Nutr. 2014, 11, 1–13. [CrossRef] [PubMed]
113. Okada, H.; Kuhn, C.; Feillet, H.; Bach, J.-F. The “hygiene hypothesis” for autoimmune and allergic diseases:
An update. Clin. Exp. Immunol. 2010, 160, 1–9. [CrossRef] [PubMed]
283
Nutrients 2018, 10, 904
114. Dabelea, D.; Mayer-Davis, E.J.; Saydah, S.; Imperatore, G.; Linder, B.; Divers, J.; Bell, R.; Badaru, A.;
Talton, J.W.; Crume, T.; et al. Prevalence of Type 1 and Type 2 Diabetes Among Children and Adolescents
From 2001 to 2009. JAMA 2014, 311, 1778. [CrossRef] [PubMed]
115. Aw, D.; Silva, A.B.; Palmer, D.B. Immunosenescence: Emerging challenges for an ageing population.
Immunology 2007, 120, 435–446. [CrossRef] [PubMed]
116. Plowden, J.; Renshaw-Hoelscher, M.; Engleman, C.; Katz, J.; Sambhara, S. Innate immunity in aging: Impact
on macrophage function. Aging Cell 2004, 3, 161–167. [CrossRef] [PubMed]
117. Lloberas, J.; Celada, A. Effect of aging on macrophage function. Exp. Gerontol. 2002, 37, 1325–1331. [CrossRef]
118. Morey, J.N.; Boggero, I.A.; Scott, A.B.; Segerstrom, S.C. Current directions in stress and human immune
function. Curr. Opin. Psychol. 2015, 5, 13–17. [CrossRef] [PubMed]
119. Organization, W.H. Mental health: Facing the challenges, building solutions. In Proceedings of the First
WHO European Ministerial Conference on Mental Health, Helsinki, Finland, 12–15 January 2005.
120. Corazon, S.; Nyed, P.; Sidenius, U.; Poulsen, D.; Stigsdotter, U. A Long-Term Follow-Up of the Efficacy
of Nature-Based Therapy for Adults Suffering from Stress-Related Illnesses on Levels of Healthcare
Consumption and Sick-Leave Absence: A Randomized Controlled Trial. Int. J. Environ. Res. Public Health
2018, 15, 137. [CrossRef] [PubMed]
121. Dhabhar, F.S. Effects of stress on immune function: The good, the bad, and the beautiful. Immunol. Res. 2014,
58, 193–210. [CrossRef] [PubMed]
122. Godbout, J.P.; Glaser, R. Stress-induced immune dysregulation: Implications for wound healing, infectious
disease and cancer. J. Neuroimmune Pharmacol. 2006, 1, 421–427. [CrossRef] [PubMed]
123. Steptoe, A.; Hamer, M.; Chida, Y. The effects of acute psychological stress on circulating inflammatory factors
in humans: A review and meta-analysis. Brain Behav. Immun. 2007, 21, 901–912. [CrossRef] [PubMed]
124. Glaser, R.; MacCallum, R.C.; Laskowski, B.F.; Malarkey, W.B.; Sheridan, J.F.; Kiecolt-Glaser, J.K. Evidence for
a shift in the Th-1 to Th-2 cytokine response associated with chronic stress and aging. J. Gerontol. A. Biol. Sci.
Med. Sci. 2001, 56, M477–M482. [CrossRef] [PubMed]
125. Veru, F.; Dancause, K.; Laplante, D.P.; King, S.; Luheshi, G. Prenatal maternal stress predicts reductions in
CD4+ lymphocytes, increases in innate-derived cytokines, and a Th2 shift in adolescents: Project Ice Storm.
Physiol. Behav. 2015, 144, 137–145. [CrossRef] [PubMed]
126. Dhabhar, F.S.; Malarkey, W.B.; Neri, E.; McEwen, B.S. Stress-induced redistribution of immune
cells—From barracks to boulevards to battlefields: A tale of three hormones—Curt Richter Award Winner.
Psychoneuroendocrinology 2012, 37, 1345–1368. [CrossRef] [PubMed]
© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
284
nutrients
Review
Obesity, Inflammation, Toll-Like Receptor 4 and
Fatty Acids
Marcelo Macedo Rogero 1,2, * and Philip C. Calder 3,4
1 Nutritional Genomics and Inflammation Laboratory, Department of Nutrition, School of Public Health,
University of São Paulo, 01246-904 São Paulo, Brazil
2 Food Research Center (FoRC), CEPID-FAPESP, Research Innovation and Dissemination Centers São Paulo
Research Foundation, São Paulo 05468-140, Brazil
3 Human Development and Health Academic Unit, Faculty of Medicine, University of Southampton,
Southampton SO16 6YD, UK; P.C.Calder@soton.ac.uk
4 National Institute for Health Research Southampton Biomedical Research Centre, University Hospital
Southampton National Health Service Foundation Trust and University of Southampton, Southampton
SO16 6YD, UK
* Correspondence: mmrogero@usp.br; Tel.: +55-11-3061-7850
Abstract: Obesity leads to an inflammatory condition that is directly involved in the etiology of
cardiovascular diseases, type 2 diabetes mellitus, and certain types of cancer. The classic inflammatory
response is an acute reaction to infections or to tissue injuries, and it tends to move towards
resolution and homeostasis. However, the inflammatory process that was observed in individuals
affected by obesity and metabolic syndrome differs from the classical inflammatory response in
certain respects. This inflammatory process manifests itself systemically and it is characterized by a
chronic low-intensity reaction. The toll-like receptor 4 (TLR4) signaling pathway is acknowledged as
one of the main triggers of the obesity-induced inflammatory response. The aim of the present review
is to describe the role that is played by the TLR4 signaling pathway in the inflammatory response
and its modulation by saturated and omega-3 polyunsaturated fatty acids. Studies indicate that
saturated fatty acids can induce inflammation by activating the TLR4 signaling pathway. Conversely,
omega-3 polyunsaturated fatty acids, such as eicosapentaenoic acid and docosahexaenoic acid,
exert anti-inflammatory actions through the attenuation of the activation of the TLR4 signaling
pathway by either lipopolysaccharides or saturated fatty acids.
1. Obesity
Obesity is a multifactorial and polygenic condition that has become a very concerning public
health issue that is affecting both developed and developing countries [1–3]. Overweight individuals
(defined as body mass index (BMI) ≥ 25 kg/m2 ) account for approximately 30% of the global
population, i.e., 2.1 billion people, of whom more than 600,000 are classified as obese (defined as
BMI ≥ 30 kg/m2 ) [4]. The analysis conducted by the Global Burden of Disease Study 2013 showed
that the overweight prevalence increased to 27.5% of adults and 47.1% of children in the past three
decades [5]. The prevalence of obesity is currently higher in developed countries; nevertheless,
approximately two-thirds of the obese population lives in developing countries [6]. Based on the
current scenario, it is estimated that up to 50% of the global population will be classified as overweight
or obese by 2030 [7]. Approximately 35% of adult individuals and 17% of children and adolescents
(2 to 19 years old) are considered to be obese (defined by values above the 95th percentile of the BMI
curve of these age groups) in the United States. It is estimated that approximately 300,000 people die
due to obesity in the United States (U.S.) every year, which is the second highest cause of preventable
death [8].
Cardiovascular diseases, type 2 diabetes (DM2), non-alcoholic fatty liver disease, and cancer
stand out among the main health issues that are responsible for morbidity related to the obesity [9].
Obesity treatment and the treatment of its associated complications in developing countries has led
to significant cost increases in healthcare. Costs that are linked to DM2, in particular, stand out,
since 20–30% of overweight people present with a DM2 diagnosis, while 85% of diabetic patients are
overweight or obese [10]. Calle et al. [11] conducted a prospective study of more than one million
men and women and found that the lowest mortality rates, for all causes, in both men and women,
occur in individuals with BMIs that are between 23.5 and 24.9 and 22.00 and 23.4 kg/m2 , respectively.
Another study including 900,000 adult individuals found that BMIs that were above 25 kg/m2 were
associated with a 30% increase in general mortality rate per each 5 kg/m2 increase [12].
Obesity results from the interactions of different factors, including genetic, metabolic, behavioral,
and environmental ones. Accordingly, the dramatic increase in obesity prevalence rates suggests
that behavioral and environmental components are the main factors that are responsible for obesity,
with an emphasis on eating habits and exercise. With regard to eating, modern societies converge to an
eating pattern called the Western diet, which is characterized by the intake of foods with high energy
densities. Such densities derive from the high contents of fat and carbohydrate, especially sugars, that
are found in these food types, a fact that contributes to obesity development [13,14].
The profile of fatty acids that are present in a diet may also be relevant to obesity. It is
worth highlighting that, according to anthropological and epidemiological studies, humans from
the Paleolithic Era—40,000 years ago—consumed a ratio of omega-6 (ω-6) to omega-3 (ω-3)
polyunsaturated fatty acids of approximately 1, mainly due to a high intake of marine and vegetable
sources of ω-3 polyunsaturated fatty acids (PUFAs). However, there was a significant increase in
the intake of lipids, trans fatty acids, and ω-6 PUFAs after the Industrial Revolution, as well as a
small increase in the intake of ω-3 fatty acids; meanwhile, intakes of vitamins C and E decreased.
Such changes are particularly relevant if one takes into account the participation of these nutrients
in the inflammatory response, which is linked to the physiopathology of different non-transmissible
chronic diseases, such as obesity, DM2, cardiovascular diseases, hypertension, and cancer [15–17].
286
Nutrients 2018, 10, 432
287
Nutrients 2018, 10, 432
plasma triglyceride concentration and promotes insulin resistance [42]. Also, in humans, BAT activity
was found to be inversely related to BMI and fat mass [43]. Furthermore, visceral adipose tissue
inflammation may also be linked to the lower BAT volume, since TNF-α has been shown to induce
brown adipocyte apoptosis and to hamper BAT differentiation [44].
Obesity is a relevant causal factor in the etiology of insulin-action resistance. Thus,
obese patients present with reduced insulin action in the skeletal muscle due to lower
phosphorylation of the tyrosine residues of the insulin receptor substrate (IRS)-1 and the reduced
phosphatidylinositol-4,5-bisphosphate 3-kinase (PI3K) activity in this tissue. Such an outcome can
cause a further reduction in insulin-induced glucose transport into the muscle tissue [45].
Figure 1. Interaction between M1 and M2 macrophages and adipocytes. Abbreviations: IL, interleukin;
MCP, monocyte chemotactic protein; NEFAs, non-esterified fatty acids; TNF, tumor necrosis factor.
288
Nutrients 2018, 10, 432
greatly integrated [50]. The excessive intake of obesity-associated nutrients can be detected by innate
recognition receptors, and this results in the activation of pro-inflammatory signaling pathways as
well as in stress responses in many parts of the body. This causes low-intensity chronic inflammation,
defined by Hotamisligil et al. [30] as metabolic inflammation or as meta-inflammation, which is
different from the classic inflammatory response. Moreover, the genesis of this inflammation is closely
related to lifestyle and mainly to the quality of diet and exercise [51].
Meta-inflammation development is associated with a wide and integrated network of intracellular
signal pathways, among which inhibitor of nuclear factor kappa-B kinase subunit beta (IKK-β) and
c-Jun N-terminal kinase 1 (JNK-1) stand out. These proteins induce the synthesis of inflammatory
mediators in different cell types. IKK-β and JNK-1 activation results in activating the transcription
factors nuclear factor kappa B (NF-κB) and the activating protein (AP)-1, which translocate to the
cell nucleus and activate the transcription of many genes encoding the proteins that are involved
in inflammation, including TNF-α and COX-2. This process allows for the continuity of the
inflammatory reaction, which is associated with conditions, such as atherogenesis and insulin-action
resistance [52,53].
This systemic inflammatory response mainly originates from adipose tissue, which produces
a wide variety of pro-inflammatory cytokines and chemokines, called adipokines [23]. However,
currently, it is known that there are other tissues involved in meta-inflammation, such as the liver [54],
pancreas [55], hypothalamus [56,57], and skeletal muscle [58]. It seems likely that the chronic low-grade
inflammation that develops in adipose tissue with obesity is “transferred” to these other tissues through
the appearance of active inflammatory mediators in the bloodstream.
In the context of inflammation and obesity, the role of gut microbiota in the development of
metabolic disease should be noted. Studies have shown that certain bacteria populations produce
enzymes that increase the efficiency of nutrient digestion, leading to an improved nutrient supply to
the host, therefore, contributing to increased energy storage in the adipose tissue. The resulting increase
in body adiposity can trigger the development of insulin resistance. There is also evidence that the gut
microbiome can modulate that genes that are involved in energy storage and expenditure [59–62].
In 2004, Backhed et al. [61] reported that conventionally reared mice had a 42% increase in body fat
and a 47% increases in periepididymal adipose tissue when compared to germ-free mice. Furthermore,
transfer of the microbiota from the bowel of the conventional mouse to the gut of the germ-free
mouse resulted in a 57% increase in body fat in two weeks, although feed consumption decreased.
This result highlights the important role that the intestinal microbiota plays in energy homeostasis
and its potential involvement in the etiology of obesity. Germ-free mice are resistant to diet -induced
adiposity, which is associated with increased activity of AMP-activated protein kinase (AMPK) in liver
and muscle and increased expression of adipose factor that is induced by fasting (Fiaf) in the small
intestine [62]. On the other hand, the inoculation of the microbiota of conventional mice fed with this
diet into germ-free animals results in an increase in adiposity [59].
It should also be noted that the dysbiosis that is associated with consuming a high-fat diet has
been shown to increase intestinal permeability, which results in a greater translocation of LPS from
the intestinal lumen to the blood circulation. This metabolic endotoxemia is associated with increased
body fat, glucose intolerance, and increased expression of proinflammatory mediators and macrophage
infiltration in white adipose tissue [60].
289
Nutrients 2018, 10, 432
Innate immune system receptors that are capable of recognizing PAMPs are called pattern
recognition receptors, and these induce the expression of pro-inflammatory cytokines—for example,
TNF-α and IL-1β—as well as activating the host’s antimicrobial defense mechanisms, such as the
synthesis of reactive oxygen and nitrogen species, including hydrogen peroxide and nitric oxide (NO),
respectively [66,67]. PAMP recognition can induce cluster of differentiation 80 (CD80) and cluster of
differentiation 86 (CD86) costimulatory molecules on the surface of cells, presenting antigens, as well
as inducing small antigenic peptides that are linked to major histocompatibility complex (MHC) class
II molecules in cell membranes that present antigens to CD4+ T lymphocytes so activating adaptive
immune responses [68].
The innate immune system recognizes PAMPs through toll-like receptors (TLRs) that are a family
of transmembrane proteins that are responsible for playing an essential role in the innate immune
system [69]. The main function of the TLR protein lies in controlling inflammatory and immunological
responses. TLRs can recognize a whole variety of microbial PAMPs. Eleven different TLRs have been
identified in humans and thirteen among all mammals [70]. TLRs belong to the IL-1 receptor (IL-1R)
superfamily, which have a significant homology in their cytoplasmic regions, such as in the Toll/IL-1R
(TIR) domain. The TIR domain is needed for the interaction and recruiting of many adaptive molecules
that are involved in the activation of signaling pathways [67].
TLRs are expressed in different cell compartments and are recognized by many PAMPs deriving
from viruses, pathogenic bacteria, fungi, and protozoa. TLR1, TLR2, TLR4, TLR5, TLR6, and TLR11
are expressed in the cellular membrane, whereas TLR3, TLR7, TLR8 and TLR9 are expressed in
intracellular compartments, such as the endosome and the endoplasmic reticulum. Based on the amino
acid sequence and on the genomic structure, TLRs can be divided into five subfamilies: TLR2, TLR3,
TLR4, TLR5, and TLR9. The subfamily TLR2 comprises TLR1, TLR2, TLR6, and TLR10, whereas the
subfamily TLR9 encompasses TLR7, TLR8, and TLR9 [71–73].
TLR4 was the first TLR reported in humans; it is expressed in innate immune cells, including
monocytes, macrophages, and dendritic cells, as well as in other cell types, like adipocytes, enterocytes,
and muscle cells. As indicated above, LPS is the primary agonist for TLR4 [74]. LPS is an integral
structural component that is found in the external membrane of Gram-negative bacteria as well as
representing one of the most powerful microbial inflammation indicators. It is a complex glycolipid
composed of one hydrophilic polysaccharide and one hydrophobic domain called lipid A [75].
There is some evidence that saturated fatty acids can also bind to TLR4 and activate TLR4-mediated
signaling pathways [76,77]. Also, there are other endogens ligands for TLR4, like heat shock protein
(Hsp) 60, Hsp 70, type III repeat extra domain A of fibronectin, oligosaccharides of hyaluronic acid,
polysaccharide fragments of heparan sulfate, and fibrinogen [78]. In the context of obesity, the increase
in the plasma fibrinogen levels, which represents a positive acute phase protein, acts as a factor that
is involved in the activation of the TLR4 pathway, and, consequently, in the amplification of the
inflammatory response [79].
The interaction between LPS and TLR4 induces the synthesis of pro-inflammatory cytokines, such
as TNF-α, IL-1β, IL-6, IL-8, and IL-12, which, in turn, work as endogenous inflammatory mediators
by interacting with receptors found in different target cells. In addition to cytokines, macrophages
release a whole variety of biological mediators in response to LPS, including platelet activation factor,
prostaglandins, enzymes, and reactive oxygen and nitrogen species, such as superoxide anion and
nitric oxide (NO). The synthesis of these pro-inflammatory mediators by monocytes and macrophages
is designed to inhibit the growth and the dissemination of pathogens and to eliminate them either
directly or through induction of adaptive immune responses [63,80].
LPS initially binds to the LPS-binding protein (LBP), which is found in the blood or in extracellular
spaces. This protein promotes LPS binding to the CD14 molecule, which, in turn, is moored to the
lipid bilayer by means of a glycophosphatidylinositol group that is found in most cells, except for
endothelial ones. CD14 can also exist as a soluble protein, and, in this case, can lead LPS to the
cell surface. The CD14 molecule is not found in transmembrane and intracellular domains; thus,
290
Nutrients 2018, 10, 432
it cannot trigger signal transduction processes on its own. When LPS binds to CD14, LBP dissociates
itself and the LPS-CD14 complex physically associates with TLR4. Such a receptor needs an additional
molecule, the so-called extracellular accessory protein (MD2), which binds to the TLR4 extracellular
complex in order to recognize LPS [71].
Following ligand binding, TLRs dimerize and undergo conformational changes that are required
for the subsequent recruitment of cytosolic TIR domain-containing adaptor molecules, including the
cytoplasmic adapter protein MyD88. The association between TLR4 and MyD88 gathers proteins
from the IL-1 receptor associated kinase (IRAK) family. Two members (IRAK4 and IRAK1) are
phosphorylated in sequence, and this disrupts them from the receptor complex and promotes their
association with TNF receptor associated factor 6 (TRAF6). TRAF6 then activates mitogen activated
protein kinase (MAPK) proteins. These kinases can activate the AP-1 transcription factor [81].
The transcription factor NF-κB, which is found in a dimeric form in the cytoplasm of
non-stimulated cells, is inactive when it is associated with κB inhibitors (IκB) (Figure 2). The family of
IκB proteins includes IκBα, IκBβ, IκBε, and Bcl-3, as well as the carboxy-terminal regions of NF-κB1
(p105) and NF-κB2 (p100). The IκB proteins bind to different NF-κB dimers, although they have
different affinities and specificities; therefore, besides the different NF-κB dimers that are found in a
specific cell type, there are a large number of combinations of the IκB and the NF-κB dimers [82,83].
Figure 2. Toll-like receptor 4 (TLR4) induced signaling activates the transcription factor NFκB.
LBP: LPS-binding protein; LPS: lipopolysaccharides; IRAK: IL-1 receptor associated kinase; TRAF6:
TNF receptor associated factor 6; MAPK: mitogen activated protein kinase; IKK: inhibitor of nuclear
factor kappa-B kinase; iNOs: inducible nitric oxide synthase.
Via MAPK, TRAF6 activates the IκB kinase complex (IKK), which is composed of two catalytic
subunits (IKKα and IKKβ) and one regulatory subunit (IKKγ), and has the capacity to induce
IκB phosphorylation. This phosphorylation results in IκB dissociation from the NF-κB complex
and its subsequent polyubiquitination, which, in turn, leads to IkB degradation (mediated by the
26S proteasome) [73,81]. This process allows for the NF-κB dimer to translocate into the nucleus and
to activate the transcription of many κB-dependent genes, such as the genes of pro-inflammatory
cytokines, including TNF-α, IL-1β, IL-6, COX-2, and inducible nitric oxide synthase (iNOS) (Figure 2).
NF-κB also stimulates the synthesis of IκB. Accordingly, the newly synthesized IκB binds to NF-κB
and suppresses its activity, providing a feedback inhibition mechanism [74,81]. There are five members
of the family of NF-κB transcription factors in mammals: NF-κB1 (p105/p50), NF-κB2 (p100/p52),
RelA (p65), RelB, and c-Rel, which can dimerize to form homodimers and heterodimers that, in turn,
291
Nutrients 2018, 10, 432
are associated with specific transcriptional responses to different stimuli. NF-κB1and NF-κB2 do
not contain transcriptional activation domains and their homodimers work as repressors. On the
other hand, Rel-A, Rel-B, and c-Rel drive the transcriptional activation domain, and, except for Rel-B,
are capable of forming homodimers and heterodimers along with other members of this family
of proteins. Consequently, the balance between different NF-κB homodimers and heterodimers
regulates the transcriptional activity level. It is worth highlighting that these proteins are expressed
in a specific cell and tissue pattern, which leads to an additional level of regulation. NF-κB1 (p50)
and RelA, for example, are broadly expressed, and, therefore, the p50/RelA heterodimer is the most
common NF-κB-binding activity inducer [82,83].
Human monocytes express TLR1, TLR2, TLR4, TLR5, TLR6, TLR8, and TLR9; but TLR2 and TLR4
are the receptors that are most commonly expressed in these cells. The expression of TLR2 and TLR4 in
the plasma membrane of monocytes has been confirmed by flow cytometry; TLR2 and TLR4-binding
(by peptidoglycan and LPS, respectively) generates pro-inflammatory cytokine secretion in these cells.
Moreover, TLR2 and TLR4 activation recruits monocytes and forms foam cells in murine models of
atherosclerosis [30,84].
Studies that were conducted in vitro with cell cultures showed the negative effects of
pro-inflammatory cytokines deriving from TLR4 signal pathway activation on glucose uptake and
on the metabolism of fatty acids [33,85,86]. TLR4 gene deletion in mice has a protective effect against
adipose tissue inflammation and against the resistance to insulin action that is induced by the intake of
a high fat diet, a fact that points towards the causal role played by TLR4 in metabolic changes driven
by over-eating and obesity [87,88].
Humans with type I diabetes exhibit a greater expression of TLR2 and TLR4 in the cellular
membrane in monocytes, as well as greater MyD88 protein content and IRAK phosphorylation in
monocytes in the peripheral blood than in control groups [89]. Individuals with DM2 show increased
cellular membrane levels of TLR2 and TLR4 in blood monocytes, as well as a higher concentration of
IL-1β, IL-6, IL-8, and TNF-α in serum than in controls [90]. Similarly, TLR2, TLR4, and MyD88 are
more highly expressed in blood mononuclear cells and in the abdominal subcutaneous white adipose
tissue in obese and diabetic individuals than in patients with normal weight [63,80]. Also, overweight
and obese people showed increased expression of TLR2 and TLR4 on peripheral blood mononuclear
cells and in adipose tissue in comparison with lean people; the expression levels of TLR2 and TLR4
increased significantly with increasing body mass index [91].
Furthermore, insulin-action resistance in obese individuals can increase the expression of TLR4,
which depends on the designated PU.1 transcription factor, which, in turn, regulates the gene
expression that is related to the activation and the differentiation of myeloid cells, including the
TLR2, TLR4, and TLR9 receptors [92,93]. Insulin has a suppressive effect on the expression of TLR4 and
on the activity of the PU.1 transcription factor; however, the suppressive effect of the hormone would
be expected to be reduced due to the insulin-action resistance related to obesity. Such a reduction
would increase the expression of TLR4 in peripheral blood monocytes [94]. In view of this, it seems
that the increase of the inflammatory response favors the occurrence of resistance to the action of the
insulin, through the activation of the IKK-β and JNK kinases that reduce the activation of IRS-1 in
the insulin signaling pathway. Conversely, the presence of insulin resistance favors the expression of
TLR4, suggesting that insulin resistance promotes inflammation.
As described earlier, the TLR4 pathway increases the expression of pro-inflammatory cytokines,
such as TNF-α, IL-1, and IL-6, by activating the transcription factors NF-κB and AP-1. These cytokines,
in turn, increase the hepatic synthesis of CRP, which is the classic positive acute phase reactant and
the most studied and accepted inflammatory biomarker. CRP is often used in clinical practice due to
its high stability (mean half-life of 19 hours) and its rapid production in response to inflammatory
stimuli [95,96]. It is important to note that other inflammatory biomarkers, such as IL-6, TNF-α,
the intercellular adhesion molecule (ICAM)-1, P-selectin, E-selectin, the monocyte chemotactic protein
292
Nutrients 2018, 10, 432
(MCP)-1, fibrinogen, and soluble CD40, have been characterized as predictors of cardiovascular disease,
regardless of other cardiovascular risk factors [19,26].
Dietary lipids can cause changes in the expression patterns of TLRs [97]. Ingestion of a high
calorie (910 kcal), high lipid (51 g), and high carbohydrate (88 g) meal by normal weight individuals
caused significant changes in TLR in the post-prandial period, with TLR2 and TLR4 increasing in blood
mononuclear cells. This reinforces the potential importance of postprandial inflammation for obesity,
DM2, and cardiovascular disease physiopathology [98,99]. A high-fat meal also leads to increased
NF-κB activation in the post-prandial period, as well as increased leucocyte activation, as assessed by
the surface expression of CD11a, CD11b, and CD62L [100], and metabolic endotoxemia (i.e., increased
plasma LPS levels) [101].
293
Nutrients 2018, 10, 432
insulin signaling pathways by inducing IRS-1 phosphorylation at serine residue position 307 [111].
This process reduces its interactions with the insulin receptor, and, consequently, diminishes the
insulin-induced signal transduction. Moreover, saturated fatty acids induce insulin-action resistance
due to the antagonistic action of the peroxisome proliferator-activated receptor-gamma coactivator
(PGC)-1 alpha. Such a process induces the expression of mitochondrial genes that are involved with
oxidative phosphorylation and with glucose capture, which is mediated by insulin [112,113].
294
Nutrients 2018, 10, 432
The ingestion of alpha-linolenic acid can also modulate the inflammatory response in humans.
For example, Caughey et al. [132] observed a significant reduction of TNF-α, IL-1β, TXB2 , and PGE2
production by LPS-stimulated mononuclear cell cultures that were obtained from healthy subjects who
consumed approximately 14 g/day alpha-linolenic acid for four weeks as compared to baseline and to
a control group. The effect of α-linolenic acid may have been mediated through its conversion to EPA.
With regard to the molecular effects of EPA and DHA on inflammatory-response modulation,
studies have shown that these fatty acids inhibit the expression of inflammatory genes, such as COX-2,
iNOS, and IL-1 in macrophages [103,108]. In contrast to the stimulating effect of saturated fatty acids
on TLR2 and TLR4 activation, EPA and DHA are capable of mitigating the activation of the NF-κB
transcription factor pathway that is induced by various agonists [103,133,134]. Thus, DHA reduces
NF-κB pathway activation and the expression of cytokines and COX-2 induced by TLR agonists,
such as lipopeptides (TLR2) and LPS (TLR4) in macrophages [89]. In addition, there is reduced gene
expression of COX-2 that is induced by LPS in monocytes from the peripheral blood of individuals
who use fish oil supplements [103,108]. The synthesis of the cytokines IL-1, IL-2, and TNF-α was also
mitigated after stimulation with LPS in vitro by mononuclear cells from the peripheral blood from
individuals that were supplemented with 18 g of fish oil per day for six weeks [135].
In addition, EPA and DHA present another mechanism to modulate the inflammatory response
by binding to G-protein coupled receptor 120 (GPR120), which is also known as free fatty acid
receptor 4 (FFA4). GPR120 activation induced by EPA or DHA leads to β-arrestin 2 recruitment to
the plasma membrane, where this protein binds to GPR120. Subsequently, the GPR120/β-arrestin
2 complex is internalized into the cytoplasmic compartment, where this complex binds to the
TAK1-binding protein (TAB1). This process impairs the association between TAB1 and the kinase
activated by the growth factor beta (TAK1), and, consequently, results in reduced TAK1 activation
and in reduced activity of the IKK-β/NF-κB and JNK/AP-1 signaling pathways. Accordingly,
the TAB1/TAK1 binding is a convergence point of stimuli that are induced by the TLR4 signaling
pathway and of the TNF receptor (TNFR). The mitigation of TAK-1 activation by DHA leads to the
reduced expression of genes with pro-inflammatory actions, such as TNF-α and IL-6 [136,137].
Other mechanisms that are related to the EPA and DHA effects concern their capacities to
bind to peroxisome proliferator activated receptors (PPARs), including the isoforms PPAR-alpha,
PPAR-gamma, and PPAR-beta/delta. PPARs are a group of nuclear receptors that are coded for by
different genes. PPAR isoforms form heterodimers with the retinoid X receptor (RXR) and bind to
peroxisome proliferator response elements (PPRE) in the region that is responsible for promoting the
target genes that are involved in lipid metabolism and in the inflammatory response; subsequently,
they modulate the expression of these genes [138]. PPAR-alpha and PPAR-gamma activations reduce
the expression of genes that code for proteins presenting pro-inflammatory actions through inhibition
of NF-κB activation. It is worth emphasizing that EPA and DHA directly interact with PPARs, and,
therefore, modulate the expression of genes that are involved in lipid metabolism and the inflammatory
response [139]. Furthermore, the anti-inflammatory effects of EPA and DHA on this signaling pathway
can occur due to diminished nicotinamide adenine dinucleotide phosphate (NADPH) oxidase activity,
which leads to lower TLR4 recruitment for lipid rafts and TLR4 dimerization [102]. Moreover, the
lower NADPH oxidase activity also decreases the production of reactive oxygen species, which, in turn,
are necessary to activate the TLR4 signaling pathway. Another possible mechanism of action of the
ω-3 fatty acids concerns the capacity of incorporating DHA into the plasma membrane, which can
lead to reduced TLR4 translocation for lipid rafts formation. This decreases TLR4 pathway activation,
and, consequently, decreases NF-κB activation [102,140,141].
Figure 3 shows the main molecular mechanisms related to the effects of saturated and omega-3
fatty acids on the TLR4 pathway.
295
Nutrients 2018, 10, 432
Figure 3. Molecular mechanism of the effects of saturated (16:0) and omega-3 polyunsaturated fatty
acids (EPA, DHA) on the TLR4 and NFkB pathways. The arrows → indicate activation and the
arrows indicate inhibition. Abbreviations: TNFα, Tumor necrosis factor; TNFR1, Tumor necrosis
factor receptor 1; LPS, Lipopolysaccharides; 16:0, palmitic acid; TLR4, Toll-like receptor 4; GPR120,
G-protein coupled receptor 120; EPA, eicosapentaenoic acid; DHA, Docosahexaenoic acid; IRS-1, Insulin
receptor substrate 1; Ser-P, phosphorylated serine residues; PPARγ, Peroxisome proliferator-activated
receptor gamma; JNK, c-Jun N-terminal kinases; IKK β, inhibitor of nuclear factor kappa-B kinase
subunit beta; IkB, NFKB Inhibitor; P, phosphate; AP-1, Activator protein 1.
5. Conclusions
The inflammatory process that occurs in obese people differs from the classical inflammatory
response in certain respects. This inflammatory process manifests itself systemically and is
characterized by a chronic low-intensity reaction. In this context, the TLR4 signaling pathway has
been recognized as one of the main triggers in increasing the obesity-induced inflammatory response.
This pathway responds to the increased exposure to saturated fatty acids and to LPS. Both of these
are relevant in the context of obesity, with saturated fatty acids arising from within the adipose
tissue triglyceride stores and the LPS arising from increased intestinal permeability perhaps due
to an altered gut microbiota. Adipose tissue driven inflammation increases insulin resistance, both
locally and systemically, so contributing to the co-morbidities of obesity, like DM2. Studies indicate
that omega-3 fatty acids, namely EPA and DHA, have an anti-inflammatory effect, which involves
attenuating the activation of the TLR4 signaling pathway. This has relevant implications for
reducing meta-inflammation, and, consequently, resistance to insulin action and the risk of DM2
and cardiovascular disease in obese individuals. The omega-3 fatty acids can oppose the action of both
classic TLR agonists (e.g., LPS) and saturated fatty acids in this regard.
Acknowledgments: The authors would like to thank The São Paulo Research Foundation and the Brazilian
National Council for Scientific and Technological Development (CNPq) for the financial support.
Author Contributions: Literature searching and initial manuscript preparation were performed by M.M.R.
The manuscript was revised and finalized by P.C.C. and M.M.R.
Conflicts of Interest: The authors declare that they have no conflict of interest.
References
1. Flegal, K.M.; Carroll, M.D.; Kit, B.K.; Ogden, C.L. Prevalence of obesity and trends in the distribution of
body mass index among US adults, 1999–2010. JAMA 2012, 307, 491–497. [CrossRef] [PubMed]
296
Nutrients 2018, 10, 432
2. Kopelman, P.G. Obesity as a medical problem. Nature 2000, 404, 635–643. [CrossRef] [PubMed]
3. Yach, D.; Stuckler, D.; Brownell, K.D. Epidemiologic and economic consequences of the global epidemics of
obesity and diabetes. Nat. Med. 2006, 12, 62–66. [CrossRef] [PubMed]
4. WHO—World Health Organization. World Health Organization Obesity and overweight Fact Sheet (2016).
Available online: http://www.who.int/mediacentre/factsheets/fs311/en/ (accessed on 30 January 2018).
5. Ng, M.; Fleming, T.; Robinson, M.; Thomson, B.; Graetz, N.; Margono, C.; Mullany, E.C.; Biryukov, S.;
Abbafati, C.; Abera, S.F.; et al. Global, regional, and national prevalence of overweight and obesity in
children and adults during 1980–2013: A systematic analysis for the Global Burden of Disease Study 2013.
Lancet 2014, 384, 766–781. [CrossRef]
6. Alexandratos, N.; Bruinsma, J. World Agriculture Towards 2030/2050: The 2012 Revision; FAO: Rome, Italy, 2012.
7. United Nations News Centre, 2015. United Nations News Centre. Available online:
http://www.un.org/sustainabledevelopment/blog/2015/07/un-projects-world-population-to-reach-8-5
-billion-by-2030-driven-by-growth-in-developing-countries/ (accessed on 17 January 2018).
8. Ogden, C.L.; Carroll, M.D.; Curtin, L.R.; Lamb, M.M.; Flegal, K.M. Prevalence of high body mass index in
US children and adolescents, 2007–2008. JAMA 2010, 303, 242–249. [CrossRef] [PubMed]
9. Mokdad, A.H.; Ford, E.S.; Bowman, B.A.; Dietz, W.H.; Vinicor, F.; Bales, V.S.; Marks, J.S. Prevalence of obesity,
diabetes, and obesity-related health risk factors, 2001. JAMA 2003, 289, 76–79. [CrossRef] [PubMed]
10. Daousi, C.; Casson, I.F.; Gill, G.V.; MacFarlane, I.A.; Wilding, J.P.; Pinkney, J.H. Prevalence of obesity in type
2 diabetes in secondary care: Association with cardiovascular risk factors. Postgrad. Med. J. 2006, 82, 280–284.
[CrossRef] [PubMed]
11. Calle, E.E.; Thun, M.J.; Petrelli, J.M.; Rodriguez, C.; Heath, C.W., Jr. Body-mass index and mortality in a
prospective cohort of U.S. adults. N. Engl. J. Med. 1999, 341, 1097–1105. [CrossRef] [PubMed]
12. Prospective Studies Collaboration; Whitlock, G.; Lewington, S.; Sherliker, P.; Clarke, R.; Emberson, J.;
Halsey, J.; Qizilbash, N.; Collins, R.; Peto, R. Body-mass index and cause-specific mortality in 900 000 adults:
Collaborative analyses of 57 prospective studies. Lancet 2009, 373, 1083–1096. [CrossRef] [PubMed]
13. Amuna, P.; Zotor, F.B. Epidemiological and nutrition transition in developing countries: Impact on human
health and development. Proc. Nutr. Soc. 2008, 67, 82–90. [CrossRef] [PubMed]
14. Vandevijvere, S.; Chow, C.C.; Hall, K.D.; Umali, E.; Swinburn, B.A. Increased food energy supply as a major
driver of the obesity epidemic: A global analysis. Bull. World Health Organ. 2015, 93, 446–456. [CrossRef]
[PubMed]
15. Roberts, C.K.; Barnard, R.J. Effects of exercise and diet on chronic disease. J. Appl. Physiol. (1985) 2005, 98, 3–30.
[CrossRef] [PubMed]
16. Simopoulos, A.P.; DiNicolantonio, J.J. The importance of a balanced ω-6 to ω-3 ratio in the prevention and
management of obesity. Open Heart 2016, 3, e000385. [CrossRef] [PubMed]
17. Galli, C.; Calder, P.C. Effects of fat and fatty acid intake on inflammatory and immune responses: A critical
review. Ann. Nutr. Metab. 2009, 55, 123–139. [CrossRef] [PubMed]
18. Calder, P.C. Polyunsaturated fatty acids and inflammatory processes: New twists in an old tale. Biochimie
2009, 91, 791–795. [CrossRef] [PubMed]
19. Calder, P.C.; Ahluwalia, N.; Albers, R.; Bosco, N.; Bourdet-Sicard, R.; Haller, D.; Holgate, S.T.; Jönsson, L.S.;
Latulippe, M.E.; Marcos, A.; et al. A consideration of biomarkers to be used for evaluation of inflammation
in human nutritional studies. Br. J. Nutr. 2013, 109, 1–34. [CrossRef] [PubMed]
20. Molfino, A.; Amabile, M.I.; Monti, M.; Muscaritoli, M. Omega-3 Polyunsaturated Fatty Acids in Critical
Illness: Anti-Inflammatory, Proresolving, or Both? Oxid. Med. Cell. Longev. 2017, 2017, 5987082. [CrossRef]
[PubMed]
21. Serhan, C.N.; Chiang, N.; Dalli, J. New pro-resolving n-3 mediators bridge resolution of infectious
inflammation to tissue regeneration. Mol. Asp. Med. 2017. [CrossRef] [PubMed]
22. Calder, P.C. Omega-3 fatty acids and inflammatory processes. Nutrients 2010, 2, 355–374. [CrossRef]
[PubMed]
23. Calder, P.C.; Ahluwalia, N.; Brouns, F.; Buetler, T.; Clement, K.; Cunningham, K.; Esposito, K.; Jönsson, L.S.;
Kolb, H.; Lansink, M.; et al. Dietary factors and low-grade inflammation in relation to overweight and
obesity. Br. J. Nutr. 2011, 106, 5–78. [CrossRef] [PubMed]
24. Calder, P.C. The role of marine omega-3 (n-3) fatty acids in inflammatory processes, atherosclerosis and
plaque stability. Mol. Nutr. Food Res. 2012, 56, 1073–1080. [CrossRef] [PubMed]
297
Nutrients 2018, 10, 432
25. Calder, P.C. Marine omega-3 fatty acids and inflammatory processes: Effects, mechanisms and clinical
relevance. Biochim. Biophys. Acta 2015, 1851, 469–484. [CrossRef] [PubMed]
26. Calder, P.C. Fatty acids and inflammation: The cutting edge between food and pharma. Eur. J. Pharmacol.
2011, 668, 50–58. [CrossRef] [PubMed]
27. Calder, P.C.; Yaqoob, P. Marine omega-3 fatty acids and coronary heart disease. Curr. Opin. Cardiol. 2012, 27,
412–419. [CrossRef] [PubMed]
28. Emilsson, V.; Thorleifsson, G.; Zhang, B.; Leonardson, A.S.; Zink, F.; Zhu, J.; Carlson, S.; Helgason, A.;
Walters, G.B.; Gunnarsdottir, S.; et al. Genetics of gene expression and its effect on disease. Nature 2008, 452,
423–428. [CrossRef] [PubMed]
29. Hotamisligil, G.S.; Shargill, N.S.; Spiegelman, B.M. Adipose expression of tumor necrosis factor-alpha: Direct
role in obesity-linked insulin resistance. Science 1993, 259, 87–91. [CrossRef] [PubMed]
30. Hotamisligil, G.S. Inflammation and metabolic disorders. Nature 2006, 444, 860–867. [CrossRef] [PubMed]
31. Wellen, K.E.; Hotamisligil, G.S. Inflammation, stress, and diabetes. J. Clin. Investig. 2005, 115, 1111–1119.
[CrossRef] [PubMed]
32. Zeyda, M.; Stulnig, T.M. Adipose tissue macrophages. Immunol. Lett. 2007, 112, 61–67. [CrossRef] [PubMed]
33. Lumeng, C.N.; Bodzin, J.L.; Saltiel, A.R. Obesity induces a phenotypic switch in adipose tissue macrophage
polarization. J. Clin. Investig. 2007, 117, 175–184. [CrossRef] [PubMed]
34. Cancello, R.; Clement, K. Is obesity an inflammatory illness? Role of low-grade inflammation and
macrophage infiltration in human white adipose tissue. BJOG 2006, 113, 1141–1147. [CrossRef] [PubMed]
35. Cave, M.C.; Hurt, R.T.; Frazier, T.H.; Matheson, P.J.; Garrison, R.N.; McClain, C.J.; McClave, S.A. Obesity,
inflammation, and the potential application of pharmaconutrition. Nutr. Clin. Pract. 2008, 23, 16–34.
[CrossRef] [PubMed]
36. Ferrante, A.W. Obesity-induced inflammation: A metabolic dialogue in the language of inflammation.
J. Intern. Med. 2007, 262, 408–414. [CrossRef] [PubMed]
37. Cannon, B.; Nedergaard, J. Brown adipose tissue: Function and physiological significance. Physiol. Rev. 2004,
84, 277–359. [CrossRef] [PubMed]
38. Nedergaard, J.; Bengtsson, T.; Cannon, B. Unexpected evidence for active brown adipose tissue in adult
humans. Am. J. Physiol. Endocrinol. Metab. 2007, 293, E444–E452. [CrossRef] [PubMed]
39. Cinti, S. The adipose organ. Prostaglandins Leukot. Essent. Fatty Acids 2005, 73, 9–15. [CrossRef] [PubMed]
40. Cao, L.; Choi, E.Y.; Liu, X.; Martin, A.; Wang, C.; Xu, X.; During, M.J. White to brown fat phenotypic switch
induced by genetic and environmental activation of a hypothalamic-adipocyte axis. Cell Metab. 2011, 14,
324–338. [CrossRef] [PubMed]
41. Lowell, B.B.; Susulic, V.; Hamann, A.; Lawitts, J.A.; Himms-Hagen, J.; Boyer, B.B.; Kozak, L.P.; Flier, J.S.
Development of obesity in transgenic mice after genetic ablation of brown adipose tissue. Nature 1993, 366,
740–742. [CrossRef] [PubMed]
42. Connolly, E.; Morrisey, R.D.; Carnie, J.A. The effect of interscapular brown adipose tissue removal on
body-weight and cold response in the mouse. Br. J. Nutr. 1982, 47, 653–658. [CrossRef] [PubMed]
43. van Marken Lichtenbelt, W.D.; Vanhommerig, J.W.; Smulders, N.M.; Drossaerts, J.M.; Kemerink, G.J.;
Bouvy, N.D.; Schrauwen, P.; Teule, G.J. Cold-activated brown adipose tissue in healthy men. N. Engl. J. Med.
2009, 360, 1500–1508. [CrossRef] [PubMed]
44. Nisoli, E.; Briscini, L.; Giordano, A.; Tonello, C.; Wiesbrock, S.M.; Uysal, K.T.; Cinti, S.; Carruba, M.O.;
Hotamisligil, G.S. Tumor necrosis factor alpha mediates apoptosis of brown adipocytes and defective brown
adipocyte function in obesity. Proc. Natl. Acad. Sci. USA 2000, 97, 8033–8038. [CrossRef] [PubMed]
45. Samuel, V.T.; Shulman, G.I. Mechanisms for insulin resistance: Common threads and missing links. Cell
2012, 148, 852–871. [CrossRef] [PubMed]
46. Guo, S. Insulin signaling, resistance, and the metabolic syndrome: Insights from mouse models into disease
mechanisms. J. Endocrinol. 2014, 220, 1–23. [CrossRef] [PubMed]
47. Hotamisligil, G.S.; Davis, R.J. Cell Signaling and Stress Responses. Cold Spring Harb. Perspect. Biol. 2016, 8.
[CrossRef] [PubMed]
48. Han, M.S.; Jung, D.Y.; Morel, C.; Lakhani, S.A.; Kim, J.K.; Flavell, R.A.; Davis, R.J. JNK expression by
macrophages promotes obesity-induced insulin resistance and inflammation. Science 2013, 339, 218–222.
[CrossRef] [PubMed]
298
Nutrients 2018, 10, 432
49. Tilg, H.; Moschen, A.R. Insulin resistance, inflammation, and non-alcoholic fatty liver disease.
Trends Endocrinol. Metab. 2008, 19, 371–379. [CrossRef] [PubMed]
50. Kirwan, A.M.; Lenighan, Y.M.; O’Reilly, M.E.; McGillicuddy, F.C.; Roche, H.M. Nutritional modulation of
metabolic inflammation. Biochem. Soc. Trans. 2017, 45, 979–985. [CrossRef] [PubMed]
51. Egger, G.; Dixon, J. Obesity and chronic disease: Always offender or often just accomplice? Br. J. Nutr. 2009,
102, 1238–1242. [CrossRef] [PubMed]
52. Ertunc, M.E.; Hotamisligil, G.S. Lipid signaling and lipotoxicity in metaflammation: Indications for metabolic
disease pathogenesis and treatment. J. Lipid Res. 2016, 57, 2099–2114. [CrossRef] [PubMed]
53. Hotamisligil, G.S. Inflammation, metaflammation and immunometabolic disorders. Nature 2017, 542,
177–185. [CrossRef] [PubMed]
54. Cai, D.; Yuan, M.; Frantz, D.F.; Melendez, P.A.; Hansen, L.; Lee, J.; Shoelson, S.E. Local and systemic insulin
resistance resulting from hepatic activation of IKK-beta and NF-kappaB. Nat. Med. 2005, 11, 183–190.
[CrossRef] [PubMed]
55. Ehses, J.A.; Perren, A.; Eppler, E.; Ribaux, P.; Pospisilik, J.A.; Maor-Cahn, R.; Gueripel, X.; Ellingsgaard, H.;
Schneider, M.K.; Biollaz, G.; et al. Increased number of islet-associated macrophages in type 2 diabetes.
Diabetes 2007, 56, 2356–2370. [CrossRef] [PubMed]
56. De Souza, C.T.; Araujo, E.P.; Bordin, S.; Ashimine, R.; Zollner, R.L.; Boschero, A.C.; Saad, M.J.; Velloso, L.A.
Consumption of a fat-rich diet activates a proinflammatory response and induces insulin resistance in the
hypothalamus. Endocrinology 2005, 146, 4192–4199. [CrossRef] [PubMed]
57. Milanski, M.; Arruda, A.P.; Coope, A.; Ignacio-Souza, L.M.; Nunez, C.E.; Roman, E.A.; Romanatto, T.;
Pascoal, L.B.; Caricilli, A.M.; Torsoni, M.A.; et al. Inhibition of hypothalamic inflammation reverses
diet-induced insulin resistance in the liver. Diabetes 2012, 61, 1455–1462. [CrossRef] [PubMed]
58. Varma, V.; Yao-Borengasser, A.; Rasouli, N.; Nolen, G.T.; Phanavanh, B.; Starks, T.; Gurley, C.; Simpson, P.;
McGehee, R.E., Jr.; Kern, P.A.; et al. Muscle inflammatory response and insulin resistance: Synergistic
interaction between macrophages and fatty acids leads to impaired insulin action. Am. J. Physiol.
Endocrinol. Metab. 2009, 296, 1300–1310. [CrossRef] [PubMed]
59. Turnbaugh, P.J.; Bäckhed, F.; Fulton, L.; Gordon, J.I. Diet-induced obesity is linked to marked but reversible
alterations in the mouse distal gut microbiome. Cell Host Microbe 2008, 3, 213–223. [CrossRef] [PubMed]
60. Cani, P.D.; Bibiloni, R.; Knauf, C.; Neyrinck, A.M.; Delzenne, N.M.; Burcelin, R. Changes in gut microbiota
control metabolic endotoxemia-induced inflammation in high-fat diet-induced obesity and diabetes in mice.
Diabetes 2008, 57, 1470–1481. [CrossRef] [PubMed]
61. Backhed, F.; Ding, H.; Wang, T.; Hooper, L.V.; Koh, G.Y.; Nagy, A.; Semenkovich, C.F.; Gordon, J.I. The gut
microbiota as an environmental factor that regulates fat storage. Proc. Natl. Acad. Sci. USA 2004, 101,
15718–15723. [CrossRef] [PubMed]
62. Bäckhed, F.; Manchester, J.K.; Semenkovich, C.F.; Gordon, J.I. Mechanisms underlying the resistance to
diet-induced obesity in germ-free mice. Proc. Natl. Acad. Sci. USA 2007, 104, 979–984. [CrossRef] [PubMed]
63. Basith, S.; Manavalan, B.; Lee, G.; Kim, S.G.; Choi, S. Toll-like receptor modulators: A patent review
(2006–2010). Expert Opin. Ther. Pat. 2011, 21, 927–944. [CrossRef] [PubMed]
64. Pandey, S.; Kawai, T.; Akira, S. Microbial sensing by Toll-like receptors and intracellular nucleic acid sensors.
Cold Spring Harb. Perspect. Biol. 2014, 7, a016246. [CrossRef] [PubMed]
65. Satoh, T.; Akira, S. Toll-Like Receptor Signaling and Its Inducible Proteins. Microbiol. Spectr. 2016, 4.
[CrossRef]
66. Beutler, B.A. TLRs and innate immunity. Blood 2009, 113, 1399–1407. [CrossRef] [PubMed]
67. Moresco, E.M.; LaVine, D.; Beutler, B. Toll-like receptors. Curr. Biol. 2011, 21, 488–493. [CrossRef] [PubMed]
68. Lin, Q.; Li, M.; Fang, D.; Fang, J.; Su, S.B. The essential roles of Toll-like receptor signaling pathways in sterile
inflammatory diseases. Int. Immunopharmacol. 2011, 11, 1422–1432. [CrossRef] [PubMed]
69. Carvalho, F.A.; Aitken, J.D.; Vijay-Kumar, M.; Gewirtz, A.T. Toll-like receptor-gut microbiota interactions:
Perturb at your own risk! Annu. Rev. Physiol. 2012, 74, 177–198. [CrossRef] [PubMed]
70. Trudler, D.; Farfara, D.; Frenkel, D. Toll-like receptors expression and signaling in glia cells in
neuro-amyloidogenic diseases: Towards future therapeutic application. Mediators Inflamm. 2010, 2010.
[CrossRef] [PubMed]
71. Connolly, D.J.; O’Neill, L.A. New developments in Toll-like receptor targeted therapeutics. Curr. Opin. Pharmacol.
2012, 12, 510–518. [CrossRef] [PubMed]
299
Nutrients 2018, 10, 432
72. Triantafilou, M.; Triantafilou, K. The dynamics of LPS recognition: Complex orchestration of multiple
receptors. J. Endotoxin Res. 2005, 11, 5–11. [CrossRef] [PubMed]
73. Beutler, B. Innate immunity: An overview. Mol. Immunol. 2004, 40, 845–859. [CrossRef] [PubMed]
74. Dobrovolskaia, M.A.; Vogel, S.N. Toll receptors, CD14, and macrophage activation and deactivation by LPS.
Microbes Infect. 2002, 4, 903–914. [CrossRef]
75. Triantafilou, M.; Triantafilou, K. Lipopolysaccharide recognition: CD14, TLRs and the LPS-activation cluster.
Trends Immunol. 2002, 23, 301–304. [CrossRef]
76. Huang, S.; Rutkowsky, J.M.; Snodgrass, R.G.; Ono-Moore, K.D.; Schneider, D.A.; Newman, J.W.; Adams, S.H.;
Hwang, D.H. Saturated fatty acids activate TLR-mediated proinflammatory signaling pathways. J. Lipid Res.
2012, 53, 2002–2013. [CrossRef] [PubMed]
77. Pal, D.; Dasgupta, S.; Kundu, R.; Maitra, S.; Das, G.; Mukhopadhyay, S.; Ray, S.; Majumdar, S.S.;
Bhattacharya, S. Fetuin-A acts as an endogenous ligand of TLR4 to promote lipid-induced insulin resistance.
Nat. Med. 2012, 18, 1279–1285. [CrossRef] [PubMed]
78. Gay, N.J.; Gangloff, M. Structure and function of Toll receptors and their ligands. Annu. Rev. Biochem. 2007,
76, 141–165. [CrossRef] [PubMed]
79. Al-ofi, E.; Coffelt, S.B.; Anumba, D.O. Fibrinogen, an endogenous ligand of Toll-like receptor 4, activates
monocytes in pre-eclamptic patients. J. Reprod. Immunol. 2014, 103, 23–28. [CrossRef] [PubMed]
80. Könner, A.C.; Brüning, J.C. Toll-like receptors: Linking inflammation to metabolism. Trends Endocrinol. Metab.
2011, 22, 16–23. [CrossRef] [PubMed]
81. Kawai, T.; Akira, S. Pathogen recognition with Toll-like receptors. Curr. Opin. Immunol. 2005, 17, 338–344.
[CrossRef] [PubMed]
82. Caamaño, J.; Hunter, C.A. NF-kappaB family of transcription factors: Central regulators of innate and
adaptive immune functions. Clin. Microbiol. Rev. 2002, 15, 414–429. [CrossRef] [PubMed]
83. Li, Q.; Verma, I.M. NF-kappaB regulation in the immune system. Nat. Rev. Immunol. 2002, 2, 725–734.
[CrossRef] [PubMed]
84. Cole, J.E.; Georgiou, E.; Monaco, C. The expression and functions of toll-like receptors in atherosclerosis.
Mediat. Inflamm. 2010, 2010, 393946. [CrossRef] [PubMed]
85. Jager, J.; Grémeaux, T.; Cormont, M.; Le Marchand-Brustel, Y.; Tanti, J.F. Interleukin-1beta-induced insulin
resistance in adipocytes through down-regulation of insulin receptor substrate-1 expression. Endocrinology
2007, 148, 241–251. [CrossRef] [PubMed]
86. Kim, F.; Pham, M.; Luttrell, I.; Bannerman, D.D.; Tupper, J.; Thaler, J.; Hawn, T.R.; Raines, E.W.;
Schwartz, M.W. Toll-like receptor-4 mediates vascular inflammation and insulin resistance in diet-induced
obesity. Circ. Res. 2007, 100, 1589–1596. [CrossRef] [PubMed]
87. Saberi, M.; Woods, N.B.; de Luca, C.; Schenk, S.; Lu, J.C.; Bandyopadhyay, G.; Verma, I.M.; Olefsky, J.M.
Hematopoietic cell-specific deletion of toll-like receptor 4 ameliorates hepatic and adipose tissue insulin
resistance in high-fat-fed mice. Cell Metab. 2009, 10, 419–429. [CrossRef] [PubMed]
88. Shi, H.; Kokoeva, M.V.; Inouye, K.; Tzameli, I.; Yin, H.; Flier, J.S. TLR4 links innate immunity and fatty
acid-induced insulin resistance. J. Clin. Investig. 2006, 116, 3015–3025. [CrossRef] [PubMed]
89. Devaraj, S.; Dasu, M.R.; Rockwood, J.; Winter, W.; Griffen, S.C.; Jialal, I. Increased toll-like receptor
(TLR) 2 and TLR4 expression in monocytes from patients with type 1 diabetes: Further evidence of a
proinflammatory state. J. Clin. Endocrinol. Metab. 2008, 93, 578–583. [CrossRef] [PubMed]
90. Ahmad, R.; Al-Mass, A.; Atizado, V.; Al-Hubail, A.; Al-Ghimlas, F.; Al-Arouj, M.; Bennakhi, A.; Dermime, S.;
Behbehani, K. Elevated expression of the toll like receptors 2 and 4 in obese individuals: Its significance for
obesity-induced inflammation. J. Inflamm. (Lond). 2012, 9, 48. [CrossRef] [PubMed]
91. Dasu, M.R.; Devaraj, S.; Park, S.; Jialal, I. Increased toll-like receptor (TLR) activation and TLR ligands in
recently diagnosed type 2 diabetic subjects. Diabetes Care 2010, 33, 861–868. [CrossRef] [PubMed]
92. Haehnel, V.; Schwarzfischer, L.; Fenton, M.J.; Rehli, M. Transcriptional regulation of the human toll-like
receptor 2 gene in monocytes and macrophages. J. Immunol. 2002, 168, 5629–5637. [CrossRef] [PubMed]
93. Rehli, M.; Poltorak, A.; Schwarzfischer, L.; Krause, S.W.; Andreesen, R.; Beutler, B. PU.1 and interferon
consensus sequence-binding protein regulate the myeloid expression of the human Toll-like receptor 4 gene.
J. Biol. Chem. 2000, 275, 9773–9781. [CrossRef] [PubMed]
300
Nutrients 2018, 10, 432
94. Ghanim, H.; Mohanty, P.; Deopurkar, R.; Sia, C.L.; Korzeniewski, K.; Abuaysheh, S.; Chaudhuri, A.;
Dandona, P. Acute modulation of toll-like receptors by insulin. Diabetes Care 2008, 31, 1827–1831. [CrossRef]
[PubMed]
95. Wang, C. Obesity, inflammation, and lung injury (OILI): The good. Mediat. Inflamm. 2014, 2014, 978463.
[CrossRef] [PubMed]
96. Wu, Y.; Potempa, L.A.; El Kebir, D.; Filep, J.G. C-reactive protein and inflammation: Conformational changes
affect function. Biol. Chem. 2015, 396, 1181–1197. [CrossRef] [PubMed]
97. Aljada, A.; Mohanty, P.; Ghanim, H.; Abdo, T.; Tripathy, D.; Chaudhuri, A.; Dandona, P. Increase in
intranuclear nuclear factor kappaB and decrease in inhibitor kappaB in mononuclear cells after a mixed
meal: Evidence for a proinflammatory effect. Am. J. Clin. Nutr. 2004, 79, 682–690. [CrossRef] [PubMed]
98. Ghanim, H.; Abuaysheh, S.; Sia, C.L.; Korzeniewski, K.; Chaudhuri, A.; Fernandez-Real, J.M.; Dandona, P.
Increase in plasma endotoxin concentrations and the expression of Toll-like receptors and suppressor of
cytokine signaling-3 in mononuclear cells after a high-fat, high-carbohydrate meal: Implications for insulin
resistance. Diabetes Care 2009, 32, 2281–2287. [CrossRef] [PubMed]
99. Burdge, G.C.; Calder, P.C. Plasma cytokine response during the postprandial period: A potential causal
process in vascular disease? Br. J. Nutr. 2005, 93, 3–9. [CrossRef] [PubMed]
100. Van Oostrom, A.J.; Rabelink, T.J.; Verseyden, C.; Sijmonsma, T.P.; Plokker, H.W.; De Jaegere, P.P.;
Cabezas, M.C. Activation of leukocytes by postprandial lipemia in healthy volunteers. Atherosclerosis
2004, 177, 175–182. [CrossRef] [PubMed]
101. Erridge, C.; Attina, T.; Spickett, C.M.; Webb, D.J. A high-fat meal induces low-grade endotoxemia: Evidence
of a novel mechanism of postprandial inflammation. Am. J. Clin. Nutr. 2007, 86, 1286–1292. [CrossRef]
[PubMed]
102. Hwang, D.H.; Kim, J.A.; Lee, J.Y. Mechanisms for the activation of Toll-like receptor 2/4 by saturated fatty
acids and inhibition by docosahexaenoic acid. Eur. J. Pharmacol. 2016, 785, 24–35. [CrossRef] [PubMed]
103. Lee, J.Y.; Sohn, K.H.; Rhee, S.H.; Hwang, D. Saturated fatty acids, but not unsaturated fatty acids, induce the
expression of cyclooxygenase-2 mediated through Toll-like receptor 4. J. Biol. Chem. 2001, 276, 16683–16689.
[CrossRef] [PubMed]
104. Hoshino, K.; Takeuchi, O.; Kawai, T.; Sanjo, H.; Ogawa, T.; Takeda, Y.; Takeda, K.; Akira, S. Cutting edge:
Toll-like receptor 4 (TLR4)-deficient mice are hyporesponsive to lipopolysaccharide: Evidence for TLR4 as
the Lps gene product. J. Immunol. 1999, 162, 3749–3752. [PubMed]
105. Mylona, E.E.; Mouktaroudi, M.; Crisan, T.O.; Makri, S.; Pistiki, A.; Georgitsi, M.; Savva, A.; Netea, M.G.;
van der Meer, J.W.; Giamarellos-Bourboulis, E.J.; et al. Enhanced interleukin-1β production of PBMCs from
patients with gout after stimulation with Toll-like receptor-2 ligands and urate crystals. Arthritis Res. Ther.
2012, 14, 158. [CrossRef] [PubMed]
106. Snodgrass, R.G.; Huang, S.; Choi, I.W.; Rutledge, J.C.; Hwang, D.H. Inflammasome-mediated secretion of
IL-1β in human monocytes through TLR2 activation; modulation by dietary fatty acids. J. Immunol. 2013,
191, 4337–4347. [CrossRef] [PubMed]
107. Lee, J.Y.; Zhao, L.; Youn, H.S.; Weatherill, A.R.; Tapping, R.; Feng, L.; Lee, W.H.; Fitzgerald, K.A.; Hwang, D.H.
Saturated fatty acid activates but polyunsaturated fatty acid inhibits Toll-like receptor 2 dimerized with
Toll-like receptor 6 or 1. J. Biol. Chem. 2004, 279, 16971–16979. [CrossRef] [PubMed]
108. Lee, J.Y.; Plakidas, A.; Lee, W.H.; Heikkinen, A.; Chanmugam, P.; Bray, G.; Hwang, D.H. Differential
modulation of Toll-like receptors by fatty acids: Preferential inhibition by n-3 polyunsaturated fatty acids.
J. Lipid Res. 2003, 44, 479–486. [CrossRef] [PubMed]
109. Caricilli, A.M.; Nascimento, P.H.; Pauli, J.R.; Tsukumo, D.M.; Velloso, L.A.; Carvalheira, J.B.; Saad, M.J.
Inhibition of toll-like receptor 2 expression improves insulin sensitivity and signaling in muscle and white
adipose tissue of mice fed a high-fat diet. J. Endocrinol. 2008, 199, 399–406. [CrossRef] [PubMed]
110. Erridge, C.; Samani, N.J. Saturated fatty acids do not directly stimulate Toll-like receptor signaling.
Arterioscler. Thromb. Vasc. Biol. 2009, 29, 1944–1949. [CrossRef] [PubMed]
111. Capurso, C.; Capurso, A. From excess adiposity to insulin resistance: The role of free fatty acids.
Vascul. Pharmacol. 2012, 57, 91–97. [CrossRef] [PubMed]
301
Nutrients 2018, 10, 432
112. Holland, W.L.; Bikman, B.T.; Wang, L.P.; Yuguang, G.; Sargent, K.M.; Bulchand, S.; Knotts, T.A.; Shui, G.;
Clegg, D.J.; Wenk, M.R.; et al. Lipid-induced insulin resistance mediated by the proinflammatory receptor
TLR4 requires saturated fatty acid-induced ceramide biosynthesis in mice. J. Clin. Investig. 2011, 121,
1858–1870. [CrossRef] [PubMed]
113. Patel, P.S.; Buras, E.D.; Balasubramanyam, A. The role of the immune system in obesity and insulin resistance.
J. Obes. 2013, 2013, 616193. [CrossRef] [PubMed]
114. Calder, P.C. n-3 Fatty acids and cardiovascular disease: Evidence explained and mechanisms explored.
Clin. Sci. (Lond). 2004, 107, 1–11. [CrossRef] [PubMed]
115. Calder, P.C.; Yaqoob, P. Understanding omega-3 polyunsaturated fatty acids. Postgrad. Med. 2009, 121,
148–157. [CrossRef] [PubMed]
116. Burghardt, P.R.; Kemmerer, E.S.; Buck, B.J.; Osetek, A.J.; Yan, C.; Koch, L.G.; Britton, S.L.; Evans, S.J. Dietary
n-3:n-6 fatty acid ratios differentially influence hormonal signature in a rodent model of metabolic syndrome
relative to healthy controls. Nutr. Metab. (Lond). 2010, 7, 53. [CrossRef] [PubMed]
117. Wan, J.B.; Huang, L.L.; Rong, R.; Tan, R.; Wang, J.; Kang, J.X. Endogenously decreasing tissue n-6/n-3 fatty
acid ratio reduces atherosclerotic lesions in apolipoprotein E-deficient mice by inhibiting systemic and
vascular inflammation. Arterioscler. Thromb. Vasc. Biol. 2010, 30, 2487–2494. [CrossRef] [PubMed]
118. Baker, E.J.; Miles, E.A.; Burdge, G.C.; Yaqoob, P.; Calder, P.C. Metabolism and functional effects of
plant-derived omega-3 fatty acids in humans. Prog. Lipid Res. 2016, 64, 30–56. [CrossRef] [PubMed]
119. Burdge, G.C.; Wootton, S.A. Conversion of alpha-linolenic acid to eicosapentaenoic, docosapentaenoic and
docosahexaenoic acids in young women. Br. J. Nutr. 2002, 88, 411–420. [CrossRef] [PubMed]
120. Goyens, P.L.; Spilker, M.E.; Zock, P.L.; Katan, M.B.; Mensink, R.P. Conversion of alpha-linolenic acid in
humans is influenced by the absolute amounts of alpha-linolenic acid and linoleic acid in the diet and not by
their ratio. Am. J. Clin. Nutr. 2006, 84, 44–53. [CrossRef] [PubMed]
121. Calder, P.C. The relationship between the fatty acid composition of immune cells and their function.
Prostaglandins Leukot. Essent. Fatty Acids 2008, 79, 101–108. [CrossRef] [PubMed]
122. Casula, M.; Soranna, D.; Catapano, A.L.; Corrao, G. Long-term effect of high dose omega-3 fatty acid
supplementation for secondary prevention of cardiovascular outcomes: A meta-analysis of randomized,
placebo controlled trials [corrected]. Atheroscler. Suppl. 2013, 14, 243–251. [CrossRef]
123. Chowdhury, R.; Warnakula, S.; Kunutsor, S.; Crowe, F.; Ward, H.A.; Johnson, L.; Franco, O.H.;
Butterworth, A.S.; Forouhi, N.G.; Thompson, S.G.; et al. Association of dietary, circulating, and supplement
fatty acids with coronary risk: A systematic review and meta-analysis. Ann. Intern. Med. 2014, 160, 398–406.
[CrossRef] [PubMed]
124. Marik, P.E.; Varon, J. Omega-3 dietary supplements and the risk of cardiovascular events: A systematic
review. Clin. Cardiol. 2009, 32, 365–372. [CrossRef] [PubMed]
125. Studer, M.; Briel, M.; Leimenstoll, B.; Glass, T.R.; Bucher, H.C. Effect of different antilipidemic agents and
diets on mortality: A systematic review. Arch. Intern. Med. 2005, 165, 725–730. [CrossRef] [PubMed]
126. Yokoyama, M.; Origasa, H.; Matsuzaki, M.; Matsuzawa, Y.; Saito, Y.; Ishikawa, Y.; Oikawa, S.;
Sasaki, J.; Hishida, H.; Itakura, H.; et al. Effects of eicosapentaenoic acid on major coronary events in
hypercholesterolaemic patients (JELIS): A randomised open-label, blinded endpoint analysis. Lancet 2007,
369, 1090–1098. [CrossRef]
127. Buettner, R.; Parhofer, K.G.; Woenckhaus, M.; Wrede, C.E.; Kunz-Schughart, L.A.; Schölmerich, J.;
Bollheimer, L.C. Defining high-fat-diet rat models: Metabolic and molecular effects of different fat types.
J. Mol. Endocrinol. 2006, 36, 485–501. [CrossRef] [PubMed]
128. Hartweg, J.; Farmer, A.J.; Holman, R.R.; Neil, A. Potential impact of omega-3 treatment on cardiovascular
disease in type 2 diabetes. Curr. Opin. Lipidol. 2009, 20, 30–38. [CrossRef] [PubMed]
129. Yu, K.; Bayona, W.; Kallen, C.B.; Harding, H.P.; Ravera, C.P.; McMahon, G.; Brown, M.; Lazar, M.A.
Differential activation of peroxisome proliferator-activated receptors by eicosanoids. J. Biol. Chem. 1995, 270,
23975–23983. [CrossRef] [PubMed]
130. Kalupahana, N.S.; Claycombe, K.; Newman, S.J.; Stewart, T.; Siriwardhana, N.; Matthan, N.;
Lichtenstein, A.H.; Moustaid-Moussa, N. Eicosapentaenoic acid prevents and reverses insulin resistance
in high-fat diet-induced obese mice via modulation of adipose tissue inflammation. J. Nutr. 2010, 140,
1915–1922. [CrossRef] [PubMed]
302
Nutrients 2018, 10, 432
131. Caughey, G.E.; Mantzioris, E.; Gibson, R.A.; Cleland, L.G.; James, M.J. The effect on human tumor necrosis
factor alpha and interleukin 1 beta production of diets enriched in n-3 fatty acids from vegetable oil or fish
oil. Am. J. Clin. Nutr. 1996, 63, 116–122. [CrossRef] [PubMed]
132. Chiang, N.; Serhan, C.N. Structural elucidation and physiologic functions of specialized pro-resolving
mediators and their receptors. Mol. Asp. Med. 2017, 58, 114–129. [CrossRef] [PubMed]
133. Sampath, H.; Ntambi, J.M. Polyunsaturated fatty acid regulation of gene expression. Nutr. Rev. 2004, 62,
333–339. [CrossRef] [PubMed]
134. Stryjecki, C.; Mutch, D.M. Fatty acid-gene interactions, adipokines and obesity. Eur. J. Clin. Nutr. 2011, 65,
285–297. [CrossRef] [PubMed]
135. Endres, S.; Meydani, S.N.; Ghorbani, R.; Schindler, R.; Dinarello, C.A. Dietary supplementation with n-3
fatty acids suppresses interleukin-2 production and mononuclear cell proliferation. J. Leukoc. Biol. 1993, 54,
599–603. [CrossRef] [PubMed]
136. Oh, D.Y.; Talukdar, S.; Bae, E.J.; Imamura, T.; Morinaga, H.; Fan, W.; Li, P.; Lu, W.J.; Watkins, S.M.; Olefsky, J.M.
GPR120 is an omega-3 fatty acid receptor mediating potent anti-inflammatory and insulin-sensitizing effects.
Cell 2010, 142, 687–698. [CrossRef] [PubMed]
137. Oh, D.Y.; Olefsky, J.M. Omega 3 fatty acids and GPR120. Cell Metab. 2012, 15, 564–565. [CrossRef] [PubMed]
138. Li, A.C.; Glass, C.K. PPAR- and LXR-dependent pathways controlling lipid metabolism and the development
of atherosclerosis. J. Lipid Res. 2004, 45, 2161–2173. [CrossRef] [PubMed]
139. Martínez-Fernández, L.; Laiglesia, L.M.; Huerta, A.E.; Martínez, J.A.; Moreno-Aliaga, M.J. Omega-3 fatty
acids and adipose tissue function in obesity and metabolic syndrome. Prostaglandins Other Lipid Mediat. 2015,
121, 24–41. [CrossRef] [PubMed]
140. Meital, L.T.; Sandow, S.L.; Calder, P.C.; Russell, F.D. Abdominal aortic aneurysm and omega-3
polyunsaturated fatty acids: Mechanisms, animal models, and potential treatment. Prostaglandins Leukot.
Essent. Fatty Acids 2017, 118, 1–9. [CrossRef] [PubMed]
141. Puglisi, M.J.; Hasty, A.H.; Saraswathi, V. The role of adipose tissue in mediating the beneficial effects of
dietary fish oil. J. Nutr. Biochem. 2011, 22, 101–108. [CrossRef] [PubMed]
© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
303
MDPI
St. Alban-Anlage 66
4052 Basel
Switzerland
Tel. +41 61 683 77 34
Fax +41 61 302 89 18
www.mdpi.com