You are on page 1of 15

Brain, Behavior, and Immunity 39 (2014) 8–22

Contents lists available at ScienceDirect

Brain, Behavior, and Immunity


journal homepage: www.elsevier.com/locate/ybrbi

Review

Aging and immunity – Impact of behavioral intervention


Ludmila Müller a,⇑, Graham Pawelec b
a
Max Planck Institute for Human Development, Lentzeallee 94, D-14195 Berlin, Germany
b
Center for Medical Research, University of Tübingen, Waldhörnlestr. 22, D-72072 Tübingen, Germany

a r t i c l e i n f o a b s t r a c t

Article history: Immune responses to pathogens to which they were not previously exposed are commonly less effective
Available online 4 December 2013 in elderly people than in young adults, whereas those to agents previously encountered and overcome in
earlier life may be amplified. This is reflected in the robust finding in many studies that the proportions
Keyword: and numbers of naïve B and T cells are lower and memory cells higher in the elderly. In addition to the
Immunosenescence ‘‘extrinsic’’ effects of pathogen exposure, ‘‘intrinsic’’ events such as age-associated differences in hae-
matopoeitic stem cells and their niches in the bone marrow associated with differences in cell maturation
and output to the periphery are also observed. In the case of T cells, the ‘‘intrinsic’’ process of thymic invo-
lution, beginning before puberty, further contributes to reducing the production of naïve T cells. Like
memory T cell populations, innate immune cells may be increased in number but decreased in efficacy
on a per-cell basis. Thus, superimposed on chronological age alone, remodelling of immunity as a result
of interactions with the environment over the life course is instrumental in shaping immune status in
later life. In addition to interactions with pathogens, host microbiome and nutrition, exercise and stress,
and many other extrinsic factors are crucial modulators of this ‘‘immunosenescence’’ process. In this
review, we briefly outline the observed immune differences between younger and older people, and dis-
cuss the possible impacts of behavioral variations thereon.
Ó 2013 Elsevier Inc. All rights reserved.

1. Introduction in sex hormone secretion patterns and their changes over the life-
span are clearly candidates intimately involved in controlling aging
The process of aging mostly reflects the biological consequences trajectories, but their exact contributions to longevity in humans
of unrepaired damage over time, and has a complex phenotype are not well-established. The modulating and reciprocal influence
associated with progressive changes in many key physiological of sex hormones on major physiological responses to environmen-
systems (Fig. 1). Aging influences an organism’s entire physiology, tal and cellular stressors, and to oxidative damage, may play a role
impacting on functions at the molecular, cellular and systemic lev- in longevity. Additional factors which are proposed to have an
els and increasing susceptibility to many major chronic diseases. influence in this context include telomere and telomerase-related
Many age-related diseases have multifactorial aetiology and aging differences, as well as changes in mitochondrial DNA (Pan and
is hypothesed to alter the highly coordinated interactions on the Chang, 2012). Although the role of gender differences in the regu-
systemic level, leading to loss of homeostasis and to decreased lation of inflammatory and regulatory pathways (such as insulin/
ability to respond to extrinsic and intrinsic challenges, resulting IGF signalling and Target of Rapamycin (TOR) signalling) is not en-
in senescence. For this reason, a comprehensive explanation of tirely elucidated, these pathways or factors clearly do play a role in
how and why we age requires an understanding of events at the longevity and aging-related diseases.
different levels of this decline. Further complexity is introduced Particularly chronic oxidative stress is known to affect those
particularly in studies on people due to the great diversity among cells constituting central regulatory systems (such as the nervous,
individuals, which results from the dynamic interaction of genet- endocrine and immune systems) and lead to disturbed communi-
ics, environment, life style, nutrition and other factors (including cation between them. This affects their functional capacity, dere-
those still not identified). One obvious example relates to differ- gulates homeostasis and thus may influence longevity (Fuente
ences between male and female gender, which influence life span Mde et al., 2011). In some way associated with these events,
for reasons that remain unclear (Nussinovitch and Shoenfeld, low-level inflammatory status is commonly found to be elevated
2012; Tower and Arbeitman, 2009). Gender-specific differences in the elderly, and is implicated in frailty and mortality. Numerous
attempts to define the role of chronic inflammation in aging have
implicated redox stress, mitochondrial damage, immunosenes-
⇑ Corresponding author. Tel.: +49 3082406380. cence, endocrinosenescence, epigenetic modifications and other
E-mail address: lmueller@mpib-berlin.mpg.de (L. Müller).

0889-1591/$ - see front matter Ó 2013 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.bbi.2013.11.015

Downloaded for dr laqif abdurrahman (laqif.abdurrahman@gmail.com) at Universitas Indonesia from ClinicalKey.com by Elsevier on May 20, 2018.
For personal use only. No other uses without permission. Copyright ©2018. Elsevier Inc. All rights reserved.
L. Müller, G. Pawelec / Brain, Behavior, and Immunity 39 (2014) 8–22 9

Body’s systems Age-related changes

Lifespan

Fig. 1. Aging alters the highly coordinated interactions on the systemic level, leading to loss of homeostasis and senescence.

phenomena. No single mechanism or theory is likely to be able to Here, we will briefly survey the cells and functions of the verte-
explain all aspects of aging – it is more likely that multiple pro- brate immune system focussing on the human immune system,
cesses contribute to this process; nonetheless, nearly all of them and the effects of aging thereon, before considering if and how
may be associated with inflammatory responses in some way (Jen- behavioral interventions might be able to restore appropriate im-
ny, 2012). mune function in the elderly.
Age-related inflammatory processes are intimately intertwined
with changes in immune function. At the same time, they are reg-
ulated by neuroendocrine hormones, including glucocorticoids, 2. Age-related changes in the immune system
dehydroepiandrosterone, and the catecholamines, epinephrine,
and norepinephrine. During the life course, age-related changes Age-related physiological changes can be very well exemplified
in endocrine function can potentially lead to the disturbance of this in the immune system, which is continuously remodelled over the
regulation. Social, occupational and psychological stressors are a life course. The most important task of the immune system is to
part of our daily life and the source of life-changing events. Chronic defend the body’s integrity against external pathogens or altered
stress of this type is also known to cause harmful effects on both internal factors, and to facilitate the maintenance of a beneficial
neuroendocrine and immune functions and may contribute (in microbiota (Pawelec, 2012). Various immune mechanisms of both
combination with age) to further increases in morbidity and mor- the innate and adaptive arms of the immune system, including dif-
tality among elderly individuals (Heffner, 2011; Hawkley and ferent cell populations, are available to respond to these challenges
Cacioppo, 2004). to bodily integrity. Recently, however, the strict distinction be-
Thus, accumulating evidence shows incontrovertibly that im- tween these two arms of immunity, classically viewed as com-
mune function changes with normal aging and independently with pletely separate, has become less obvious. It has been shown that
increased stress; and that chronic stress deregulates multiple com- innate immune cells can demonstrate memory characteristics un-
ponents of innate and adaptive immunity, leading to what might der certain conditions, and reciprocally that adaptive immune cells
be construed as premature aging of the immune system (Gouin can express receptors characteristic of innate cells, especially at
et al., 2008). At the same time it is likely that the immune system late stages of differentiation, i.e. commonly present in increased
impacts on the rate of organismal aging (Fuente Mde et al., 2011). numbers later in life and with the appearance of age-associated,
Consistent with a central role of the immune system in this pro- potentially senescent, changes (Lanier and Sun, 2009).
cess, several lifestyle strategies such as intervening to provide an The changes in the immune system that accompany human
adequate diet, physical exercise, physical and mental activity, also aging are very complex and are generally referred to as immunose-
result in improved immune functions, decreasing oxidative stress, nescence. Aging does not necessarily lead to unavoidable deterio-
and potentially increasing individual longevity. rations in immune functions, but always affects their modulation.
Finally, human beings may be considered as ‘metaorganisms’ as While many aspects of immune function decline with aging, some
a result of a close symbiotic relationship with the intestinal micro- of them remain stable, and some become overactive. Hence, dys-
biota (and indeed also systemic microbiota, such as persistent regulation rather than deterioration per se is likely to be the major
viruses). This assumption enforces an even more holistic view of explanantion for immunosenescence. Nevertheless, increased sus-
the aging process where dynamics of the interaction between envi- ceptibility to infections and autoimmune diseases, neoplasias, met-
ronment, intestinal microbiota and all physiological processes of abolic diseases, osteoporosis and neurological disorders are all in
the host must be taken into consideration (Biagi et al., 2012). some way likely to be at least partly caused by immunosenescence.

Downloaded for dr laqif abdurrahman (laqif.abdurrahman@gmail.com) at Universitas Indonesia from ClinicalKey.com by Elsevier on May 20, 2018.
For personal use only. No other uses without permission. Copyright ©2018. Elsevier Inc. All rights reserved.
10 L. Müller, G. Pawelec / Brain, Behavior, and Immunity 39 (2014) 8–22

Macrophage Neutrophil
Reduced expression Reduced chemotaxis
and function of TLR of neutrophils
Pathogens
Reduced MHC-expres- Impaired signal
sion and signaling Blood vessel transduction
DC Monocyte
Reduced chemotaxis Inflammatory Decreased superoxide
and phagocytosis proteins production

Decreased cytokine
production Cytokins Reduced apoptosis
TC
Impaired distribution
and migratory capacity Macrophage Reduced expression of
of DC TLR and MHC
Dendritic cell Monocyte
Ageing Ageing

Fig. 2. Age-related alteration in macrophages, monocytes, neutrophils and dendritic cells. TLR: toll-like receptor; MHC: major histocompatibility complex; DC: dendritic cells.

The term immunosenescence is now well-established in the sci- Nevertheless, a number of impairments in their function (such as
entific literature and includes a number of profound changes that chemotaxis, phagocytosis, production of free radicals as well as
affect both the innate and adaptive arms of the immune system. susceptibility to apoptosis) in aged individuals have been reported
The factors and mechanisms of immune senescence are multiple (Shaw et al., 2010). The expression and function of surface recep-
and include, among others, defects in the bone marrow (reflected tors recognizing pathogen-associated molecular patterns, such as
in alterations in the proportions of lymphoid and myeloid cells ex- Toll-like receptors, as well as cytokine production, and expression
ported to the periphery), thymus involution (decreasing the gener- of other surface molecules like MHC class II antigens is also re-
ation of new T cells) and intrinsic defects in formation, maturation, duced (Fulop et al., 2004). Thus, impairments in these initial de-
homeostasis and migration of peripheral lymphocytes (Gruver fences against infections could be very important in the elderly.
et al., 2007). The genetic background of aged individuals, epige-
netic changes occurring during the life course, impaired interaction 3.2. Monocytes and macrophages
of innate and adaptive immune responses, continuous reshaping of
the immune repertoire by persistent antigenic challenges, chronic These cells also play multiple important roles in the immune
low-grade inflammation and deregulation of hormonal pathways – system including key functions in phagocytosis and possibly anti-
all these factors might synergise, leading to the age-associated de- gen presentation (Fig. 2). Monocytes are very efficient in the elim-
cline of immunity commonly seen in the elderly (Jenny, 2012). ination of bacterial infections, which occurs with high prevalence
Age-dependent epigenetic changes in DNA methylation do occur in the aged population. Macrophages are directly involved in the
in cells of the immune system in the same way that overall geno- initiation of inflammatory responses, elimination of pathogens
mic methylcytosine levels decrease in other major tissues. More and tumor cells, and regulation of the adaptive immune response
demethylation is known to occur in the brain, heart, liver, small through the process of antigen presentation. They can destroy their
intestine mucosa, spleen, and also in T lymphocytes in the elderly. targets directly, or they can act indirectly through release of im-
These age-dependent changes in DNA methylation are likely to mune mediators (such as IL-1, TNF-a, and IFN-c) that activate
contribute in some way to T cell senescence (Calvanese et al., other inflammatory cells (Derhovanessian et al., 2008; Keller,
2009; Gonzalo, 2010). Additional to these intrinsic factors, also 1993). The number of blood monocytes does not appear to change
extrinsic factors, like health-related behaviors and chronic stress with advancing age, but a decreased percentage of macrophages in
exposure (Bauer, 2005) can accelerate immunosenescence and the bone marrow of aged individuals 80–100 years old has been re-
the process of aging as discussed later in this chapter. ported, which can be explained by increased apoptosis and re-
duced cellularity. Expression of MHC-II molecules in
macrophages is impaired with aging, and epigenetic mechanisms
3. Cells of the innate immune system and impact of aging
have been invoked in this decrease. The phagocytic and killing
capacities of macrophages are also reduced with age, accompanied
Innate immune responses initially call adaptive immune re-
by lower production of reactive oxygen intermediates such as NO2
sponses into play and both arms act together to eliminate patho-
and H2O2, and decreased levels of macrophage-derived cytokines
gens. Cells of the innate immune system, such as natural killer
(TNF-a and IL-1), the expression of which is modulated by epige-
cells, macrophages, dendritic cells, and neutrophils, generate a
netic mechanisms (Fernandez-Morera et al., 2010; Gonzalo, 2010).
more rapid but less finely antigen-specific immune response than
the cells of the adaptive immune system.
3.3. Dendritic cells

3.1. Neutrophils Dendritic cells (DCs) play crucial roles in the immune response
and constitute an essential bridge between the innate and adaptive
Neutrophils are the most abundant population of circulating immune systems. They are responsible for recognizing pathogen-
leukocytes and represent the first line of defence against most associated molecules, for taking up and processing pathogen anti-
types of pathogens, such as bacteria, yeast and fungi (Fig. 2, right). gens, and for presenting these antigens to T cells. They represent
These cells have a very high turnover and relatively short lifespan. professional antigen-presenting cells, without which adaptive

Downloaded for dr laqif abdurrahman (laqif.abdurrahman@gmail.com) at Universitas Indonesia from ClinicalKey.com by Elsevier on May 20, 2018.
For personal use only. No other uses without permission. Copyright ©2018. Elsevier Inc. All rights reserved.
L. Müller, G. Pawelec / Brain, Behavior, and Immunity 39 (2014) 8–22 11

Compensatory NK cell NKT cell / iNKT cell


Increase in NKT cell
increase in cell
numbers
numbers
Virus-infected
NKT cell
or modified cell
Reduced per-cell Decrease in
cytotoxicity TCR frequency of iNKT

Decreased signal IL-2 Reduced population


transduction CD161 doubling in iNKT
ARL
IFN-γ ARL
AR TNF-γ
Reduced response to IFN-γ Impaired cytokine
cytokines TNF-γ production
Perforin
Granzyme
Impaired cytokine and Shift in cytokine
chemokine production profile: Th1 to Th2
NK cell Virus-infected
Ageing or tumour cell
Ageing

Fig. 3. Age-related changes in NK- and NKT-cells. NK: natural killer cells; NKT: natural killer T cells; iNKT: invariant natural killer T cells; IL: interleukin; IFN: interferon; TNF:
tumor necrosis factor; TCR: T-cell receptor; CD: cluster of differentiation; AR: activation receptor; ARL: activation receptor ligand; Th: T-helper.

immune responses cannot take place. Moreover, the manner in 3.5. Natural killer T cells
which the DCs present antigen and costimulate T cells influences
the type of T cell response initiated, and thus the quality and inten- Natural killer T (NKT) cells share attributes of both the innate
sity of adaptive immune responses (Müller et al., 2013). This im- and adaptive arms of the immune system. They are a unique and
plies that age-associated decreased DC function leads relatively rare (ca. 0.1%) subset of CD3-positive T cells that co-
unavoidably to reduced T cell activation and proliferation. express TCRs and CD161 (NKR-P1A), undergo maturation in the
DCs seem to be affected by aging in terms of their distribution thymus and/or at extrathymic sites, and play an important role
and migratory capacity, in their antigen-processing ability as well in viral and antitumor cytolytic activity (Fig. 3, right). A subset of
as in costimulatory signal expression and cytokine production so-called invariant NKT (iNKT) cells is characterized by a TCR with
(Fig. 2, left). Although antigen presentation by DCs seems to be an invariant alpha chain. These two subsets of lymphocytes share
only subtly different in the elderly, fewer DC are found in periph- some attributes, such as expression of CD16, CD56 and CD161 with
eral blood and follicles, implying reduced efficiency of this line of NK-cells, but they are able to recognize antigens in both an
defence against infections and tumors (Gonzalo, 2010; Della Bella MHC-restricted (peptides), as well as a CD1-restricted (lipids and
et al., 2007; Derhovanessian et al., 2008). Similar to the results ob- glycolipids) manner.
tained from animal studies, the numbers of DCs, their distribution, It has been reported that absolute numbers of NKT-cells in-
and potentially their generation and development from hemato- crease with advancing age (Mahbub et al., 2011), whereas for the
poietic precursors in vivo are markedly reduced in elderly humans iNKT-cells reduced frequency and numbers has been found in aged
(Adema, 2009). individuals (DelaRosa et al., 2002) with no marked change in their
phenotypes. It was also shown that iNKT cells underwent fewer
population doublings compared to cells from young individuals
3.4. Natural killer cells (Peralbo et al., 2006), suggesting that they were already ‘‘older’’.
With aging, the cytokine profile secreted by NKT-cells also
Natural killer (NK) cells are cytotoxic lymphocytes which are in- changes. Recent studies indicate that cytokine profile of iNKT-cells
volved in early defence, recognizing virus-infected and virally or from old individuals showed a shift from a T helper (Th)-1 toward a
non-virally modified tumor cells in an MHC-unrestricted manner Th-2 pattern compared with iNKT-cells from young individuals
(Fig. 3, left). This outstanding competence of NK cells probably (Mahbub et al., 2011). Although NKT-cells are known for their abil-
makes them particularly important for cancer immune surveil- ity to influence immunological functions of APCs and T cells, only a
lance during aging. Studies on healthy elderly individuals and cen- few studies have examined the role of NKT-cells in
tenarians demonstrate that the overall NK cell number tends to immunosenescence.
increase with age, but their function on a per-cell basis decreases
and NK cells are also more likely to have a mature phenotype.
NK cells from the elderly also produce lower levels of cytokines 4. The adaptive immune system and immunosenescence
and chemokines (such as RANTES, MIP1a, and IL-8). Thus, the over-
all increase in number of NK cells can be considered as a compen- The cells of the adaptive immune system (T- and B-lympho-
satory mechanism to maintain an important level of functionality. cytes) act in a highly antigen-specific manner and imbue the sys-
An age-related impairment in perforin secretion, associated with tem with immunological memory (Medzhitov and Janeway,
defective polarization of lytic granules toward the immunological 1997) as well as regulating immune homeostasis. The receptors
synapse was recently proposed to be a reason for the age-associ- of these lymphocytes are generated through somatic recombina-
ated reduction in NK-cytotoxicity (Hazeldine et al., 2012). NK-cell tion of segments of their encoding sequences (Bonilla and Oettgen,
production of IL-2, IFN-c, TNF-a and IL-12 is also diminished in el- 2010) and in this way provide an extremely diverse repertoire of
derly people and this may contribute, among other factors, to im- receptor specificities capable of recognizing components of essen-
mune deficits associated with advanced age (Dewan et al., 2012). tially all potential pathogens. Adaptive immunity comprises a

Downloaded for dr laqif abdurrahman (laqif.abdurrahman@gmail.com) at Universitas Indonesia from ClinicalKey.com by Elsevier on May 20, 2018.
For personal use only. No other uses without permission. Copyright ©2018. Elsevier Inc. All rights reserved.
12 L. Müller, G. Pawelec / Brain, Behavior, and Immunity 39 (2014) 8–22

highly regulated interplay between innate cells like DCs or other shift in the proportions of these HSC subsets. Such changes in the
antigen-presenting cells, and T and B lymphocytes, which activate lymphoid and myeloid lineage are believed by some investigators
and promote pathogen-specific immunologic effector pathways to be central to the decline of immune competence and predispo-
and generate immunologic memory. During the first encounter sition to myeloproliferative diseases in the elderly (Beerman et al.,
with a pathogen, sets of long-lived memory T and B cells are 2010; Dykstra and de Haan, 2008).
established. In subsequent encounters with the same pathogen,
the memory cells are quickly activated to yield a more rapid and
robust protective response (Cooper, 2010). 5.2. Thymic involution

Another major feature contributing to immunosenescence is


5. Development of lymphocytes and impact of aging thymic involution (Fig. 4, right). During this process a progressive
age-related reduction in the size of the thymus takes place by
Development of lymphocytes occurs in the specialized environ- replacement of lymphoid tissue by fatty tissue and by reducing
ments of the bone marrow and thymus. The lymphoid precursors the active areas of thymopoiesis (Aspinall et al., 2010). This process
of T- and B-cells originate from the pluripotent hematopoietic stem was shown to be mediated by the upregulation of thymosuppres-
cells in the bone marrow. sive cytokines, such as IL-6, oncostatin M and leukaemia inhibitory
factor (Sempowski et al., 2000) and parallel decreases in IL-7 pro-
5.1. Immunosenescence and the haematopoietic stem cell duction, known to be important for thymopoiesis. These changes
compartment lead to the consequent decrease in the numbers of thymic epithe-
lial cells and impairment in thymopoiesis. The reduction of thymic
The haematopoietic stem cell (HSC) compartment, which is the output and therefore diminished replacement of circulating naïve T
source of the continuous replenishment of blood and the immune cells as they respond to pathogens or other antigens is thought to
system throughout life (Fig. 4, left), giving rise to all cellular (lym- be a major contributory event in the development of immunose-
phoid and myeloid) components, is negatively modulated and nescence (Aspinall et al., 2010).
functionally affected by aging (Geiger et al., 2013). The stromal ma- Normally, T cell precursors migrate from the bone marrow to
trix of the aging bone marrow, which normally nurtures and drives the thymus (Fig. 4), where they proliferate and differentiate into
stem cell production, also exhibits structural changes in terms of immature double-negative CD4 CD8 thymocytes (Schwarz and
the numbers of stromal cells and diminished IL-7 production. The Bhandoola, 2006). During T-cell development, rearrangement of
bone marrow haematopoietic compartment changes with age and the antigen receptor genes takes place, so that the cells now ex-
is increasingly composed of fatty adipose tissue (Compston, 2002; press the T-cell receptor (TCR) on their surfaces. Chronic thymus
Gruver et al., 2007). Age-related changes within the HSC compart- involution is associated with a reduced efficiency of T-cell develop-
ment may be partly dependent on ‘‘intrinsic’’ cellular aging of HSCs ment and with the reduced migration of naïve T cells in the periph-
themselves. Accumulation of DNA damage, telomere attrition and ery. The loss or great reduction of thymic function, although
epigenetic deregulation, combined with an increase in intracellular occurring early in life and therefore a developmental rather than
reactive oxygen species characterizes aged HSCs (Dykstra and de senescence process, results in contraction of the T cell repertoire
Haan, 2008; Warren and Rossi, 2009). These events of genomic and contributes to the increased incidence of infections and poten-
instability may also lead to malignant transformation of HSCs tially also cancer and autoimmune disease (Lynch et al., 2009).
and to myeloproliferative disorders. Age-associated decreased Changes in thymus activity may also be influenced by environmen-
homing efficiency, and a myeloid skewing of differentiation poten- tal conditions (prenatally and during the early period of life),
tial contribute to changes in the cellular composition of the HSC genetics, sexual dimorphism (males and females show different
compartment. The hematopoietic stem cell pool contains a hetero- patterns of thymic involution) and the thymic stroma (Gui et al.,
geneous mixture of HSCs, and aging is associated with a marked 2012). Sauce and colleagues showed that thymectomy during

Bone marrow Pluripotent HSC Young thymus Aged thymus


Aging
Development?
Decrease of HSC Reduced active area
compartment of thymopoiesis
Pre-B cell DNT
Transformation of Pro-B cell Up-regulation of
HSC and shift to DPT thymus-suppressive
myeloid lineage cytokines
CD8+T cell
Decreased frequency Reduced population
of pro-B cells of naÔve T cells

IgG IgM
Altered Ig-specificity, Oligoclonal TCR
-isotype and -idiotype repertoire

CD4+T cell Aggregation of


Reduced primary and exhausted memory
secondary response T cells
IgG
Ageing Memory B cell
Ageing
Memory CD4+T cell Memory CD8+T cell

Fig. 4. Development of lymphocytes and impact of aging. HSC: haemotopoietic stem cell; IgG: immunoglobulin G; IgM: immunoglobulin M; DNT: doppel-negative
thymocyte; DPT: doppel-positive thymocyte; CD: cluster of differentiation.

Downloaded for dr laqif abdurrahman (laqif.abdurrahman@gmail.com) at Universitas Indonesia from ClinicalKey.com by Elsevier on May 20, 2018.
For personal use only. No other uses without permission. Copyright ©2018. Elsevier Inc. All rights reserved.
L. Müller, G. Pawelec / Brain, Behavior, and Immunity 39 (2014) 8–22 13

Regular physical activity


and exercise

Reduced energy intake

Adipocyte
Practice of “healthy
behaviour” - Treg cell - Acvated
T cell
- M1-Macrophage - M2-Macrophage
- pro-inflammatory adipokines - an-inflammatory adipokines

Fig. 5. Impact of behavioral intervention on immunosenescence.

childhood leads to premature onset of many of these age associ- infections, exposure to chemicals or hormones, etc., which influ-
ated changes in the immune system (Appay et al., 2010; Sauce ence their development. These factors can alter the epigenetic pro-
and Appay, 2011), underlining the importance of vigorous naïve file of cells, and therefore have a direct effect on gene expression
T cell output in early life to build a reserve for later use. profiles and on remodelling processes. Senescent immune remod-
elling is a significant contributing factor to the increased risk and
5.3. Age-related changes of T cells severity of infections in the elderly (Dewan et al., 2012; Larbi
et al., 2008; Solana et al., 2006; Fernandez-Morera et al., 2010).
Circulating CD4+ T cells act mostly as T-helper cells, by trigger-
ing signals of B-cell activation for antibody production. A counter- 5.4. Changes in T-cell populations and impact of Cytomegalovirus
balancing subset of the CD4 population consists of regulatory T (CMV)
cells (Treg), which exert inhibitory activities on other T cells, sup-
pressing excessive or misguided immune responses. The number of As a consequence of decreased thymopoiesis with advancing
peripheral regulatory T cells can increase during aging as well as age and other age-related modulating factors, a shift in the ratio
under conditions of chronic stress (Bauer et al., 2009). The process of naïve to memory T cells takes place, in order to maintain T-cell
of immune aging seems to represent a remodelling of adaptive im- homeostasis in the periphery. Additionally, repetitive antigen
mune responses by progressive reduction of the TCR repertoires in exposure contributes to the continuous reshaping of the immune
the CD4+ and CD8+ subsets contributing to highly restricted oli- repertoire by persistent antigenic challenge, modulating the T-cell
goclonal TCR repertoire in the periphery of the elderly (Kohler pool and contributing to immunosenescence in older adults (Paw-
et al., 2005). This could lead to age-dependent increase in suscep- elec et al., 2009; Lang et al., 2013). Nearly 25% of the total CD8 T
tibility to infectious, autoimmune, malignant diseases and de- cells in CMV-positive elderly persons may be specific for a single
creased efficiency of vaccination. CMV-immunodominant epitope (Pawelec et al., 2009); this CMV-
At the single-cell level, T cells from aged individuals demon- driven expansion of CD8-positive T cells is accompanied by the loss
strate reduced functional potency compared to cells of the same of the costimulatory CD28 molecule, the process being considered
phenotype from young individuals. This reduced functional po- as a key predictor of immune incompetence in older individuals
tency of T cells was shown to be restored under some conditions (Müller et al., 2013). The age-dependent accumulation of ex-
by the addition of IL-15 (Liu et al., 2002). Normally, after their re- hausted memory T cells, releasing the pro-inflammatory cytokines
lease from the thymus, naïve T-cells may be activated by binding of CRP, IL-6, TNF-a and IFN-c, together with components and factors
their TCR to the appropriate MHC-peptide complex presented by of the innate immune system, is thought to contribute to the low-
APCs, such as dendritic cells. The crucial initial step in this T-cell grade inflammation (inflamm-aging) commonly observed in el-
activation is the reorganisation of the plasma membrane, in which derly people (Franceschi et al., 2007; Bennett et al., 2012). How-
membrane rafts cluster to the T cell/APC site. This process of lipid ever, the hypothesis that CMV is a significant driver in
raft polarization is impaired in T cells from older people. It was inflammaging has been challenged in a recent cohort study (Bart-
shown that the higher cholesterol content in these aged cells lett et al., 2012) showing that levels of pro-inflammatory and
may influence the motility of their rafts and thereby may reduce anti-inflammatory cytokines were changing equally in CMV-sero-
recruitment of signalling molecules following activation (Larbi positive and CMV-seronegative subjects over time, suggesting that
et al., 2011). CMV might be not a primary causative factor in inflammaging.
Naïve T cells from elderly people exhibit impaired differentia- The increased presence of exhausted CD28 T cells in elderly
tion into effector cells following antigen stimulation, as well as people, together with other parameters, such as an altered CD4/
other functional defects such as reduced cytokine production, as CD8 ratio and CMV-seropositivity, has led to the definition of the
well as a restricted TCR repertoire (Ferrando-Martinez et al., so-called ‘‘immune risk phenotype’’ predicting a higher 2-year
2011; Weiskopf et al., 2009). Epigenetic age-associated alterations, mortality in a longitudinal study of octa- and nonagenarians (Paw-
such as methylation of cytokine gene promoters can also contrib- elec et al., 2009). The accumulation of later-stage CD8 T cells char-
ute to perturbed immune function (Calvanese et al., 2009; Gonzalo, acterizing the IRP was not seen in people not infected with CMV,
2010). Both immune-specific genes (such as surface molecules or even when they were infected with other persistent herpesviruses
the cytokines IL-2 and IFN-c) and genes involved in general cell (Derhovanessian et al., 2011). However, the process of immunose-
homeostasis pathways, show altered DNA methylation profiles nescence may also be amplified in the context of both chronologi-
during aging (Fernandez-Morera et al., 2010; Shanley et al., 2009). cal aging and HIV-infection; thus, also during chronic HIV disease
Appropriate differentiation and activation of T cells partly de- the proportion of senescent CD8 T cells increases progressively
pends on the integrity of the genes encoding key factors, which with age and often consists of oligoclonal populations (Deeks
may be associated with an environmental component, viral et al., 2012). These cells gain suppressive functions and may also

Downloaded for dr laqif abdurrahman (laqif.abdurrahman@gmail.com) at Universitas Indonesia from ClinicalKey.com by Elsevier on May 20, 2018.
For personal use only. No other uses without permission. Copyright ©2018. Elsevier Inc. All rights reserved.
14 L. Müller, G. Pawelec / Brain, Behavior, and Immunity 39 (2014) 8–22

contribute to carcinogenesis. Interestingly, even in HIV disease, a tory markers (Hajat et al., 2010; Nazmi et al., 2010). It has also
high proportion of late-differentiated, senescent, memory CD8 T been demonstrated that psychosocial stress may be responsible
cells are nonetheless specific for CMV antigens (Dock and Effros, for accelerated telomere attrition, with shorter leukocyte telomere
2011). lengths in persons with lower socioeconomic status (Shiels et al.,
2011). Stress has been shown to be associated with increased risk
6. Inflammaging of frailty and cardiovascular diseases, apparently as result of ele-
vated levels of inflammatory markers (von Kanel et al., 2006; Jen-
As mentioned above, chronic low-grade inflammation has been ny, 2012).
repeatedly identified in seemingly healthy individuals during aging Frailty is a common syndrome in the elderly and is character-
and is characterized by increased levels of circulating pro-inflam- ized by poor mobility, weakness and low tolerance to psychologi-
matory cytokines, such as TNF-a, IL1Ra, IL-6 and markers of cal or/and physiological stress, and results from accumulation of
inflammation such as C-reactive protein (CRP). It was postulated multiple functional declines in various systems. Frailty may be
that this so-called ‘‘inflammaging‘‘ process appears to be a key considered as a kind of exhaustion of physiological reserves of
phenomenon associated with different age-related diseases and the organism, leading to decreased overall physical function, which
their pathological features (Franceschi et al., 2000, 2007). Inflam- can be characterized by weight loss (primarily at the cost of muscle
maging seems to be a universal phenomenon accompanying aging, loss), decline in muscle function and their reduced strength due to
and is associated with frailty, morbidity and mortality in elderly sarcopenia. As mentioned before, inflammatory status seems to
individuals (Jenny, 2012). Thus, age-related pro-inflammatory sta- play a decisive role in the frailty that is associated with higher
tus can be associated with patho-physiological conditions. How- prevalence of inflammatory diseases (Chang et al., 2012).
ever, it is often difficult to identify whether the inflammation is
the causative agent for these co-morbidities or the consequence 8.1. Impact of physical inactivity on inflammatory status
(Rymkiewicz et al., 2012).
An inactive life style and sedentary behavior promote the accu-
mulation of visceral fat and lead to obesity. Accumulating evidence
7. Humoral immunity indicates that obesity causes chronic low-grade inflammation and
development of systemic metabolic dysfunction that appears to be
Humoral immunity in an aged organism is known to be both aetiologically associated with obesity-linked disorders. Adipose
qualitatively and quantitatively different than in the young. There tissue acts as a key endocrine organ by releasing bioactive sub-
is a lower frequency and absolute number of pro-B lymphocytes in stances, known as adipokines, that have pro-inflammatory or
the bone marrow, along with a reduction in their ability to differen- anti-inflammatory activities (Ouchi et al., 2011). The production
tiate into pre-B lymphocytes (Fig. 4, left). Immunoglobulin diversity of pro-inflammatory adipokines in expanding fat tissue, such as
and affinity are reduced in the elderly because of impaired somatic TNF, leptin, retinol-binding protein 4, lipocalin 2, IL-6, IL-18 and
hypermutation inside the germinal center (GC), which is the main angiopoietin-like protein 2 increases, while the concentrations of
site of B cell proliferation and maturation. The generation of long- anti-inflammatory cytokines are reduced (Ouchi et al., 2011). This
term immunoglobulin-producing B lymphocytes is also impaired. process is accompanied by infiltration of adipose tissue with pro-
The numbers of functional immunoglobulin-secreting B cells and ti- inflammatory immune cells and induction of a low-grade inflam-
ters of antigen-specific immunoglobulins are decreased. B-cell sub- matory state (Fig. 5, left), which is characterized by elevated levels
sets, and thus the antibody repertoire, are altered in specificity and of circulating inflammation markers, such as IL-6, TNF and CRP.
isotype. The result is a shortened duration of humoral response Adipose tissue is infiltrated with macrophages in two separate
and reduced B-cell ability to establish specific primary and second- polarization states: M1, which produce pro-inflammatory cyto-
ary responses in elderly people (Colonna-Romano et al., 2008; kines and M2, producing anti-inflammatory cytokines. Therefore,
Ademokun et al., 2010). This process is also dependent on cognate it has been proposed that in adipose tissue a phenotypic switch
interactions between antigen-activated B cells and CD4+ T cells in- takes place toward macrophages of M1-phenotype, promoting
side the GCs. Therefore, age-related deficits in T cells can modulate the inflammatory state. This low-grade systemic inflammation is
cognate interactions between B and T cells and in this way reduce known to be associated with development of atherosclerosis, neu-
antibody avidity (Dewan et al., 2012). Aging-associated changes in rodegeneration, insulin-resistance and promotion of tumor growth
epigenetic status might also be responsible for altered B-cell func- (Gleeson et al., 2011). While these phenomena are not limited to
tion in elderly people, because of modifications in the differentiation the elderly, they often tend to be exacerbated in older people
pathways and in regulation and rearrangement of BCR/Ig genes who for one reason or another exercise less than the young.
(Shanley et al., 2009; Colonna-Romano et al., 2008). Taken together,
one can conclude that the quality of the humoral immune response 8.2. Impact of nutrition on immunosenescence
declines with progressing age.
Aging represents a major nutritional challenge, not only con-
8. Impact of life style factors on immunosenescence cerning the dietary supply of certain nutrients but also in terms
of their altered metabolism (Duncan and Flint, 2013). Micronutri-
Many lifestyle factors are known or suspected to contribute to ent deficiencies, which are very common in the elderly, have been
perceived deleterious age-associated changes to immunity. These found to be associated with a physiological decline in various body
include psychosocial parameters, stress responsiveness, physical functions, which can lead to a higher morbidity and mortality.
inactivity, macro- and micro-nutrition, all of which may play Among other micronutrients, zinc has an essential significance to
important roles in immunosenescence. health; its deficiency is responsible for various diseases. Zinc is
Stress may play a crucial role in mediating inflammation and one of the most important trace elements in the organism, with
age-associated impairments of immunity. Stress may be induced three major biological roles, as catalyst, structural, and regulatory
by a variety of factors, including situations both physical and men- ion. It plays a critical role in organism homeostasis, in immune
tal, such as caregiving or low socioeconomic status. The latter in function, in oxidative stress, in apoptosis and other areas (Chasapis
the elderly, depending on previous education and low income, et al., 2012). Thus, zinc deficiency may influence progression of
has been shown to be associated with higher levels of inflamma- many chronic diseases, including atherosclerosis, neurological

Downloaded for dr laqif abdurrahman (laqif.abdurrahman@gmail.com) at Universitas Indonesia from ClinicalKey.com by Elsevier on May 20, 2018.
For personal use only. No other uses without permission. Copyright ©2018. Elsevier Inc. All rights reserved.
L. Müller, G. Pawelec / Brain, Behavior, and Immunity 39 (2014) 8–22 15

disorders, autoimmune diseases, age-related degenerative dis- people. Some studies have shown that moderate exercise training
eases, various malignancies, and adversely affect immunological (5 days a week for 6 months) can improve Th-cell mediated im-
status, increase oxidative stress, and lead to the generation of mune functions associated with up-regulation of CD28 expression
inflammatory cytokines (Chasapis et al., 2012). Zinc deficiency is on the surface and by improving the Th1/Th2 balance (Shimizu
known to decrease innate immunity. Particularly, zinc deficiency et al., 2008). The training sessions (2 days a week for 3 months),
impairs the lytic activity of NK cells, reduces NKT cell cytotoxicity consisting of stretching and endurance exercise as well as resis-
and immune signalling, influences the neuroendocrine immune tance training using an exercise machine, led to increased numbers
pathway, and alters cytokine generation in mast cells (Mocchegiani of CD28+CD8+ cells as well as CD80+CD14+ cells in the periphery
et al., 2003; Muzzioli et al., 2009). (Shimizu et al., 2011), indicating that such training sessions might
The studies published thus far describe a critical role for nutri- upregulate monocytes and dendritic cells, thereby possibly
tion in maintaining the immune response of the aged, but they also improving T-cell mediated immunity in elderly people.
indicate the need for a more in-depth, holistic approach to deter- In recent years, the role of exercise in modulating immune re-
mining the optimal nutritional strategies that would maintain a sponses has been examined using models that may have clinical
healthy immune system in elderly people and promote their resis- relevance, such as the response to vaccines and novel antigens
tance to infection and other immune-related diseases (Pae et al., (Kohut and Senchina, 2004; Kohut et al., 2002). It has been shown,
2012). Therefore, we will need a more holistic view of the aging for example, that regular exercise is associated with improved im-
process where dynamics of the interaction between environment, mune responsiveness to influenza vaccination in older adults.
intestinal microbiota and host must be taken into consideration. Exercise-related increased antibody titre, T-cell function, macro-
We should take into account also the age-associated physiological phage response, improvement of the Th1/Th2 cytokine balance,
changes in the gastrointestinal tract, as well as age-related modifi- the level of pro-inflammatory cytokines, and changes in naïve/
cations in lifestyle, in nutritional behavior, and impaired function- memory cell ratio have also been reported (Kohut and Senchina,
ality in the immune system of the elderly population. 2004). An 10-month interventional trial with 144 participants
(aged 60–83 years) has demonstrated that cardiovascular, but not
flexibility and balance exercise intervention, resulted in a signifi-
9. Impact of behavioral interventions on immunosenescence
cant increase in seroprotection 24 weeks after vaccination, provid-
ing improved protection throughout the entire influenza season
It is now clear that a variety of genetic and environmental fac-
and reduced respiratory tract infections (Woods et al., 2009).
tors impact upon health in old age, including effects on immunity.
Thus, accumulating data suggest that exercise may be a power-
However, the relative contribution of these factors to immunose-
ful approach to restoring immune function in old populations.
nescence will have to be more accurately established. These vari-
However, more rigorous standardisation of procedures is required
ables clearly include nutrition (micro and macro) and obesity, as
to use them in the prevention of disease and in the modulation of
well as gender and ethnicity, genetic background, psychosocial
the immune system in order to reduce disease prevalence (Walsh
parameters (including stress), mental wellbeing, socioeconomic
et al., 2011; Hong, 2011).
status, early life events and exposures, physical activity, and differ-
ent chronic infections. It will be necessary to build up a solid evi-
9.2. Impact of exercise on senescent and virus-specific T lymphocytes
dence base in order to develop effective personalized
interventions, both lifestyle and pharmacologic, to extend health
As already mentioned above, aging is characterized by the accu-
span. Few current studies are attempting to meet this challenge,
mulation of late-stage, possibly terminally differentiated T cells, of
but according to the findings of a recent study on more than
which at least some may be truly senescent (mostly as a conse-
23,000 adults, a healthy lifestyle alone lowered the risk in develop-
quence of persistent viral infection), coupled with greatly reduced
ing chronic diseases with known inflammatory aetiology by 78%
or absent entry of naïve T cells into the periphery (as a conse-
(Ford et al., 2009). Thus, many chronic diseases can be prevented
quence of thymic atrophy). This leads to a drastically shrinking
by changing lifestyle and behavioral habits, particularly dietary
naïve T-cell repertoire. However, beneficial effects of the manipu-
habits, and exercise. For example, a positive effect of 3 months of
lation of certain life style factors, especially the benefits of regular
a regimen of comprehensive lifestyle changes (moderate exercise,
physical exercise, on markers of immunosenescence have been
plant-based diet, stress management and improved social support)
repeatedly reported, although the mechanisms underlying these
on increased telomerase activity was demonstrated in men with
potentially positive effects on immunity are not very well under-
low-risk prostate cancer (Ornish et al., 2008). After 5 years fol-
stood (Simpson and Guy, 2010).
low-up, relative telomere lengths of lymphocytes had increased
It has been suggested that habitual physical exercise can bene-
in the lifestyle intervention group and decreased in the control
ficially modulate immunosenescence in two ways: first, by a pre-
group. Although such changes are thought to be beneficial, more
ventive mechanism (restricting the opportunity for latent viral
investigations are needed to confirm whether this is really the case
reactivations), and second, by a restorative mechanism (Simpson,
and the full biological implications remain to be determined in
2011). For the latter process to take place, three distinct phases
large randomised, controlled trials (Ornish et al., 2013).
were proposed: (i) a selective mobilization of senescent T cells
from peripheral organs into the blood, occurring during exercise;
9.1. Effect of physical exercise on the immune system (ii) extravasation of senescent T cells from the circulation and their
enhanced apoptosis in peripheral tissue, occurring during exercise
Evidence that long-term behavioral changes, including reduced recovery; (iii) subsequent generation of naïve T cells to replace the
energy intake together with increased physical activity, may pre- deleted senescent cells. However given thymic involution, the abil-
vent, improve or even reverse age-related impairments in immune ity to regenerate naïve cells in most elderly people will be limited,
function, continues to accumulate. Lifestyle factors, such as exer- perhaps only to homeostatic proliferation of pre-existing naïve
cise and diet have been established as playing an important role cells.
in immunosenescence, and the practice of ‘‘healthy’’ behavior Acute exercise normally induces a biphasic change in circulat-
may minimize the age-associated decline of immune function ing lymphocyte numbers, with their rapid mobilization into the
(Fig. 5). Several interventions, including different types of exer- blood at the beginning of exercise (lymphocytosis) and falling be-
cises, have been proposed to restore immune function in elderly low the resting values during the early phase of exercise recovery

Downloaded for dr laqif abdurrahman (laqif.abdurrahman@gmail.com) at Universitas Indonesia from ClinicalKey.com by Elsevier on May 20, 2018.
For personal use only. No other uses without permission. Copyright ©2018. Elsevier Inc. All rights reserved.
16 L. Müller, G. Pawelec / Brain, Behavior, and Immunity 39 (2014) 8–22

(lymphocytopenia) (Simpson et al., 2008; Turner et al., 2010). By subpopulations than the chronological age (Spielmann et al.,
mechanisms not very well understood, in particular the lympho- 2011). Thus, more experimental work is required, to determine
cytes with high cytotoxic capacity are preferentially mobilized into the mechanisms playing a role in enhanced immunity associated
the circulation. It has been shown that these cells consist of highly with regular exercise (Walsh et al., 2011).
differentiated effector-memory T-cell subsets, expressing surface
markers that may be shared with senescent cells, such as KLRG1 9.3. Anti-inflammatory effects of physical exercise
and CD57 (Simpson et al., 2007).
High-intensity exercise was found to mobilise higher numbers Gleeson et al. (2011) reviewed three possible mechanisms for
of late-differentiated effector memory cells in comparison to the anti-inflammatory effects of exercise: (i) reduction of visceral
low-intensity exercise (Campbell et al., 2009). The same subpopu- fat mass, (ii) increased production and release of anti-inflamma-
lation of lymphocytes has been demonstrated to egress from the tory cytokines, so called myokines, from contracted skeletal mus-
circulation during exercise recovery (Simpson et al., 2008). Fur- cles and (iii) reduced expression of TLRs on monocytes and
thermore, it was shown that in CMV-seropositive young individu- macrophages. A reduction in the numbers of circulating pro-
als, mobilization and subsequent egress of total CD8+ lymphocytes inflammatory-type monocytes and an increase in the numbers of
was nearly twice as great as in CMV-seronegative subjects, indicat- circulating regulatory T cells has been found in the peripheral
ing that exercise may induce the mobilization of viral-specific CD8+ blood of individuals following exercise (Timmerman et al., 2008;
T cells (Turner et al., 2010). The same type of senescent and effec- Yeh et al., 2006). It is possible that chronic exercise can attenuate
tor memory cells are believed to overcrowd the immune space dur- inflammation in adipose tissue by reducing macrophage infiltra-
ing aging and persistent CMV-infection. Therefore, it was tion and by phenotypic switching from pro-inflammatory M1- to
postulated that mobilization of senescent viral-specific T-cells to- anti-inflammatory M2-type macrophages (Fig. 5). Recent data sug-
gether with frequent T-cell shifts in response to exercise may be gest that contracting skeletal muscles are able to release myokines,
the first step toward a restoration of the naïve T-cell repertoire such as IL-6, IL-8, IL-15, BDNF and others, which may act in a hor-
(Simpson et al., 2012). mone-like fashion, generating specific endocrine effects on visceral
In support of this suggestion, it has been demonstrated that fat or mediating direct anti-inflammatory effects. Some of them
lymphocytes mobilized by acute bouts of exercise are more sensi- may locally influence signalling pathways involved in fat oxidation
tive to in vitro apoptosis induced by physiological concentrations of (Pedersen, 2011).
pro-oxidant H2O2. The most apoptosis-susceptible subpopulation During exercise an exponential rise in muscle-derived IL-6 lev-
was found to be KLRG1+ and CD57+ with subpopulations express- els takes place, which seems to be responsible for a subsequent
ing CD28, CD62L, or CD11a least susceptible (Wang and Lin, elevation in circulating concentrations of the anti-inflammatory
2010). The subpopulation of CD45RA/CCR7-negative effector mem- cytokines IL-10 and IL-1 receptor antagonist, IL-1R, as well as tran-
ory T cells has been demonstrated to be more sensitive to H2O2-in- sient increased release of cortisol (Steensberg et al., 2003). Unlike
duced apoptosis than CD45RA/CCR7-positive naïve T cells macrophage-derived IL-6, the muscle-produced IL-6 mediates
(Takahashi et al., 2005), indicating that senescent T cells might anti-inflammatory functions (Pedersen, 2011). Transient increases
be less resistant to apoptosis induction. Kruger et al. (2009)have in levels of circulating cortisol and adrenalin during exercise is
demonstrated and quantified lymphocyte trafficking and death in dependent on activation of the HPA-axis and the sympathetic ner-
peripheral tissues of mice after exercise. They found that exer- vous system. Impulses from the motor centers in the brain to-
cise-induced apoptosis is a systemic phenomenon and is not lim- gether with afferent impulses from working muscles seem to
ited to the peripheral blood. It occurs in various lymphoid and evoke a rise in the sympathoadrenal activity in an intensity-depen-
non-lymphoid tissues, such as lung, spleen, bone marrow, lymph dent manner. Therefore, increases in levels of cortisol (which is a
nodes and Peyer’s patches and is dependent on the type of exercise potent anti-inflammatory regulator) and adrenalin (which down-
and its intensity. The apoptosis-inducing mechanisms, as well as regulates production of IL-1b and TNF) in plasma, are related to
kinetics, have a tissue-specific character. intensity and duration of exercise (Simpson et al., 2008). It is also
The transient periods of lymphocytopenia caused by exercise known that prolonged exercise results in decreased TLR expression
normally last 6–24 h, and peripheral blood lymphocyte counts with a subsequent deregulation of downstream inflammatory cas-
are normally restored to baseline within that period of time. Inter- cades (Stewart et al., 2005).
leukin-7, which is known to play a crucial role in homeostasis of Taken together, one might conclude that physical activity, such
naïve T cells, has been shown to be released from active skeletal as regular exercise, activates the release of hormones, myokines
muscles (Haugen et al., 2010), possibly contributing to increased and cytokines, as well as modulating expression of various im-
thymopoiesis and exercise-induced immune enhancement (Simp- mune-reactive molecules, which all contribute to anti-inflamma-
son et al., 2012). tory effects and possible attenuation of immunosenescence. The
The impact of maximal aerobic capacity as a measure of aerobic reduction of visceral fat mass alone already leads to a decreased
fitness on the age-related accumulation of potentially senescent T production and release of pro-inflammatory adipokines from fat
cells and on the proportion of naïve T cells has been examined in tissue (Fig. 5).
another recent study (Spielmann et al., 2011). This showed that
the proportion of senescent CD4+ and CD8+ T cells increased with 9.4. Caloric restriction
advancing age at the rate of ca. 10% per decade, accompanied by
a proportional reduction of CD4+ and CD8+ naïve T cells. Further- At least in the context of adiposity and inflammation mentioned
more, it was demonstrated that participants with higher aerobic above, as well as via multiple other postulated physiological ef-
fitness possessed less senescent CD4+ and CD8+ T cells and had fects, caloric restriction (CR) in humans might have beneficial ef-
more naïve CD8+ T cells, than those with low aerobic fitness. This fects in terms of lowering metabolism, reduction of visceral fat
association was stable even after adjusting for age, body mass in- and weight loss. CR has been shown to delay signs of immunose-
dex and percentage of body fat. Very interesting and important nescence in animals and is considered today as the only known
was the finding that the association between age and the method to prolong median as well as maximal lifespan in several
proportion of senescent cells disappeared when adjusted for scores tested species, from invertebrates to rodents and even including
of high aerobic fitness, demonstrating that the aerobic fitness non-human primates (Arnold et al., 2011; Anderson and Weind-
may be a stronger determinant of phenotypic shift of T-cell ruch, 2012). Thus, in rodents and non-human primates CR leads

Downloaded for dr laqif abdurrahman (laqif.abdurrahman@gmail.com) at Universitas Indonesia from ClinicalKey.com by Elsevier on May 20, 2018.
For personal use only. No other uses without permission. Copyright ©2018. Elsevier Inc. All rights reserved.
L. Müller, G. Pawelec / Brain, Behavior, and Immunity 39 (2014) 8–22 17

to attenuation of the age-related shift from naïve to memory-phe- Nutritional intervention including micronutrients and vitamins
notype T cells and maintains a higher number of naïve T cells in has also been recognized as a practical, cost-effective approach to
aged animals (Nikolich-Zugich and Messaoudi, 2005). Further- attenuating age-associated declines in immune function, vaccina-
more, the age-associated rise of pro-inflammatory cytokines, such tion efficiency, and resistance to infectious and neoplastic diseases.
as IL-6, IFN-c, and TNF-a, and the resulting pro-inflammatory state The importance of micronutrients and vitamins in proper immune
of an aged immune system can be reversed by CR. It has been re- functioning is clear, and among them, zinc is an essential element
ported that the age-related decrease in proliferative capacity of T the significance of which to health is undisputable, as discussed
cells (due to the shift from naïve to memory-phenotype T cells) above. It is therefore important that status of zinc is assessed in
can be reversed by CR (Arnold et al., 2011). Applying caloric restric- any case and zinc deficiency is corrected, since the unique proper-
tion in rodents resulted in 50% increases in lifespan compared to ties of zinc may have significant therapeutic benefits in these dis-
animals which have been ‘‘free-choice’’ feeded (Li et al., 2011). eases. It has been shown that zinc supplementation enhances
Nevertheless, there are still many open questions concerning CR, innate immunity by increasing phagocytosis, and T-cell functional-
which has been shown to be effective at improving the immune re- ity (Sheikh et al., 2010). Zinc supplementation was able to improve
sponse in unchallenged animals, although it might compromise the the generation of NK cells from CD34+ cell progenitors through in-
host’s defence against pathogenic infection and result in higher creased expression of GATA-3 transcription factor (Muzzioli et al.,
morbidity and mortality outside of the lab. Moreover, CR has been 2009). It was demonstrated that addition of zinc could modulate T
shown to delay immunosenescence in animals, but this effect cell-dependent immune reactions. For example, zinc supplementa-
needs to be verified in humans. Furthermore, short-term CR may tion to PBMC produces T cell activation through an indirect effect
well be feasible, whereas dietary restriction long-term might have that is mediated by cytokine production by other immune cells;
detrimental psychological and other effects in humans, making its but it has also been shown that higher concentrations of zinc can
practicality questionable (Jolly, 2007). Nevertheless, some indirect also directly suppress T cell function (Wellinghausen et al., 1997).
evidence for a positive effect of prolonged caloric restriction in hu- Vitamin E supplementation has been demonstrated to improve
mans has been found in the population of Okinawa, Japan, reported immune responsiveness in healthy elderly individuals by enhance-
to consume fewer calories than the mainland Japanese population ment of cell-mediated immunity (Meydani et al., 1990). It was also
and with a larger proportion of centenarians (Anderson and able to reverse many of the T-cell age-associated defects, including
Weindruch, 2012). More extensive studies are needed, some of reduced levels of phosphorylation of critical signalling proteins as
which are currently underway, investigating effects of the CR in well as to improve defective immune synapse formation. Vitamin E
humans (Rickman et al., 2011). also enhances IL-2 production, expression of several cell cycle con-
trol proteins, and T-cell proliferation (Molano and Meydani, 2012).
Intake of vitamin E above recommended levels has been shown to
9.5. Other nutritional interventions and their impact on enhance T cell function in aged animals and humans. This effect is
Immunosenescence believed to contribute to improved resistance to influenza infec-
tion and reduced incidence of upper respiratory infection in elderly
Optimization of nutrition is one of the first and theoretically people (Pae et al., 2012).
easiest and cheapest strategies that can be employed to preserve Taken together, studies have shown that nutritional interven-
health during aging. The development of ‘elderly-specific’ func- tion with the above-mentioned supplements, as well as many oth-
tional foods, containing probiotics and/or prebiotics, may help in ers being tested, may be a promising approach to restoring at least
preventing the age-related disruption of the gut environment (Gui- some elements of impaired immune function and countering
goz et al., 2008). In fact, it has been demonstrated that the mainte- diminished resistance to infection with aging. However, some lim-
nance of a ‘healthy’ gut microbiota during aging could help to delay itations are reviewed by Pae et al. concerning some nutritional reg-
or prevent the inflamm-aging process (Biagi et al., 2012). iments. For example, zinc deficiency, common in the elderly can be
Immunostimulatory properties, such as modulation of cytokine amended by zinc supplementation. However, taken in higher doses
production or adjuvant effects on T lymphocytes and NK activity, than recommended, it may adversely affect immune function. Pro-
have been demonstrated for various health-promoting Lactobacil- biotics are increasingly being demonstrated as a powerful, im-
lus and Bifidobacterium strains (Biagi et al., 2010; Blum et al., mune-modulating nutritional factor, but, to be effective, they
2002; Meydani and Ha, 2000). It has been shown that 3- and 6- require an adequate supplementation period. Moreover, their ef-
week interventions in elderly people can have positive effects, such fects seem to be strain-specific and among some strains, a syner-
as measurable increases of NK cells, tumoricidal activity and gistic effect is observed. Thus, the more in-depth holistic
monocyte phagocytic capacity (Takeda and Okumura, 2007). A pro- approach is needed for determination the optimal nutritional strat-
biotic yoghurt supplementation tested on elderly persons affected egies that would maintain a healthy immune system in the elderly
intestinal bacterial overgrowth. This intervention was able to nor- and promote their resistance to infection and other immune-re-
malize the response to endotoxin and modulate inflammatory lated diseases (Pae et al., 2012). Recent advances in this area sug-
markers in blood phagocytes (Schiffrin et al., 2010) as well as gest means of improving the outcomes of some of these
decreasing the incidence of infections in this group of elderly peo- approaches. A recent seminal advance has been the realization that
ple. Furthermore, a significant increase of phagocytosis, NK-cell probiotic supplementation is likely to require ‘‘tailor-made’’ mix-
activity and production of IL-10 were also demonstrated following tures of different bacteria in order to achieve the desired effect.
probiotic supplementation in the elderly, as well as reduced pro- Thus, Atarashi et al. (2013) entified 17 strains of bacteria preferen-
duction of the pro-inflammatory cytokines IL-6, IL-1b and TNF-a tially inducing colonic regulatory T cells (Tregs) and which reduced
(Guigoz et al., 2008; Vulevic et al., 2008). Boge et al. showed in symptoms in models of colitis and allergic diarrhea in mice.
two clinical trials that the consumption of a probiotic drink (con- Whereas single strains or even mixtures of less than the 17 was
taining Lactobacillus casei) by elderly subjects for several weeks be- not or not as effective as all 17 together, supplementation with tai-
fore and after influenza vaccination led to a significant increase in lor-made mixtures of probiotics may be effective at modulating
the influenza-specific antibody titre, demonstrating the potential immune responses where use of single strains is ineffective. This
of probiotics in improving the protective efficacy of vaccination may be one reason why the results of the many human probiotics
in the elderly population, in which it is usually considerably re- supplementation trials have often been equivocal, since mostly a
duced (Boge et al., 2009; Goodwin et al., 2006). single strain was used, even if it was the ‘‘right’’ strain. Atarashi

Downloaded for dr laqif abdurrahman (laqif.abdurrahman@gmail.com) at Universitas Indonesia from ClinicalKey.com by Elsevier on May 20, 2018.
For personal use only. No other uses without permission. Copyright ©2018. Elsevier Inc. All rights reserved.
18 L. Müller, G. Pawelec / Brain, Behavior, and Immunity 39 (2014) 8–22

et al. found that of all the bacteria in a human fecal sample, only a numbers of memory and regulatory T cells, diminished cell-medi-
mixture of 17 Clostridia strains was effective at modulating immu- ated immunity, restricted TCR-repertoire, elevated serum pro-
nity in germ-free mice. The mechanism of action was via the induc- inflammatory markers have been demonstrated to be present dur-
tion of CD4+ CD25+ FOXP3+ Tregs by bacterial antigens at the same ing chronic stress or/and glucocorticoid-rich conditions (Ashwell
time as stimulating intestinal epithelial cells to produce TGF-ß1 et al., 2000; Damjanovic et al., 2007; Effros et al., 2005; Elenkov
but not IL 6 and TNF via bacterial production of short-chain fatty and Chrousos, 1999; Franceschi et al., 2000; Globerson and Effros,
acids. Fecal samples from patients with inflammatory bowel dis- 2000; Kiecolt-Glaser et al., 2003; Sauce and Appay, 2011; Trzon-
ease and atopy have been reported to contain relatively low kowski et al., 2006; Wack et al., 1998).
amounts of some of the same bacterial strains, suggesting their rel- It is known that aging is correlated with activation of the hypo-
evance to human disease as well as in the animal model. thalamic–pituitary–adrenal (HPA) axis. Healthy older people have
an approximately twofold higher cortisol level and reciprocally to
9.6. Impact of nutraceuticals on immune system the same extent lower dehydroepiandrosterone (DHEA) level rela-
tive to young adults (Luz et al., 2003; Bauer et al., 2009). DHEA acts
In the context of supporting the ‘‘right’’ balance of gut bacteria, as an endogenous glucocorticoid antagonist, and the lack of appro-
reversing our nutritional habits ‘‘back to mother Nature’’ might priate levels of DHEA could lead to higher cortisol/DHEA ratio,
represent an additional solution for modulation of age-related combined with the consequent deregulation of key allostatic sys-
inflammatory status and associated chronic diseases. The accumu- tems in aged individuals. The HPA axis plays a critical role in the
lating data from in vitro-, animal- and clinical studies provide evi- homeostasis of the immune system; therefore a neuroendocrine
dence that greater consumption of fruits, cereals, vegetables, imbalance of the HPA axis leads to negative immunological conse-
legumes and spices is associated with a lower risk of many diseases quences. Due to the fact that all leukocytes express receptors for
(Adzersen et al., 2003; Ames and Wakimoto, 2002; Chadalapaka the neuroendocrine molecules of the HPA sympathetic-adrenal
et al., 2008; Kwak et al., 2010; Prasad et al., 2012; Zhang et al., medullary axes, it seems plausible that an increased cortisol/DHEA
2010). Consumption of diets consisting of natural foods can also ratio contributes to the immunological changes observed during
raise the amounts of plant-based nutraceuticals in the body, such aging. Indeed, similar changes have also been seen during chronic
as antioxidants and anti-inflammatory agents. These nutraceuti- psychological stress and in the course of in vitro and in vivo treat-
cals are known to cope with free radicals and to modulate the ment with glucocorticoids (Butcher et al., 2005; Khanfer et al.,
inflammatory signalling pathways in cells suppressing the onset 2011; Maninger et al., 2009). Thus, detrimental additive effects of
of age-related chronic conditions (Aggarwal and Shishodia, 2006; immunosenescence along with chronic stress may accelerate
Prasad et al., 2012). organismal aging and development of stress/aging-related pathol-
Here, we can provide only a few examples, limited to spices and ogies. Moreover, amplified neuroinflammation negatively influ-
fruits and their bioactive components, known to modulate differ- ences several aspects of neural plasticity (e.g., neurogenesis,
ent stages of tumorigenesis, including tumor cell survival, prolifer- long-term potentiation, and dendritic morphology) that can con-
ation, invasion and angiogenesis. The anticancer activities of spices tribute to the severity of neurological disorders, such as prolonged
are mediated primarily through suppression of inflammation. Bio- sickness, cognitive impairment and depressive-like complications
active components of spices, such as eugenol prevent the release of (Corona et al., 2012). Geriatric depression often occurs in persons
TNF-a and IL-b (Kim et al., 2003) (6)-gingerol blocks production of exposed to chronic stress – a state, which accelerates the geriatric
TNF-a in in vitro stimulated macrophages (Tripathi et al., 2007). depression and triggers pro-inflammatory processes with atten-
Curcumin has been shown to suppress the inflammatory mediators dant amplified immunosenescence. Geriatric depression is known
NF-jb and COX-2 (Shishodia et al., 2003), anethole inhibits NF-jb to exacerbate the symptoms of comorbid pathologies and disorders
activation and cytokine production (Chainy et al., 2000) and cinna- (Morimoto and Alexopoulos, 2011). Thus, the immune system and
maldehyde blocks age-related activation of NF-jb and targets the neuroendocrine system appear to modulate each other, pro-
inflammatory COX-2 and induces nitric oxide synthase (Kim moting the synthesis of pro-inflammatory cytokines, overproduc-
et al., 2007). Ursolicacid, which is present in many fruits, including tion of which, in turn, negatively influences behavior. On the
apple, pear, plum, bearberry, loquat, jamun and rosemary, has been contrary, a positive psychological situation has been associated
found to exert antitumor activity against colon cancer (Andersson with better heath, involving reduced activation of neuroendocrine,
et al., 2003), breast cancer (Es-saady et al., 1996), non-small cell as well as immune and inflammatory pathways. Positive wellbeing
lung cancer (Hsu et al., 2004), pancreatic cancer (Chadalapaka improved several immunological biomarkers, including not only
et al., 2008), melanoma (Harmand et al., 2005), multiple myeloma positive changes in the numbers of the immune cells, but also their
(Pathak et al., 2007), cervical cancer (Yim et al., 2006) and prostate improved functionality. Increased cellular immune competence
cancer (Zhang et al., 2010). Also several other nutraceuticals have has been linked with positive affect, including stronger NK-cell
been shown to exert anti-inflammatory and antitumor activities. cytotoxicity and increased secretory IgA responses to antigenic
More clinical studies are needed to determine amounts of the die- challenge. Positive affect has also been associated with increased
tary agents needed to delay aging and age-related diseases, and to numbers of helper T cells. There is also consistent evidence for
investigate their effects on different age groups (Prasad et al., the beneficial influence of positive affect on antibody titers and
2012). antibody responses to antigens, as well as on levels of such impor-
tant cytokines as IL-2, IL-3, IL-6 and TNFa (Dockray and Steptoe,
9.7. Chronic stress and aged immune system 2010).

External factors such as chronic psychological stress are 9.8. Stress-reducing interventions and their influence on immunity
thought to accelerate the aging process of the immune system, pre-
sumably because stress hormones are immunosuppressive. The Stress-buffering psychosocial interventions have accordingly
immunological changes, normally detected during aging, are also been shown to be effective in attenuating stress and in promoting
found to be affected to a similar degree, following chronic stress a better neuroendocrine balance in the elderly. By reducing stress
or prolonged glucocorticoid exposure in younger individuals (Bau- levels and supporting healthy behaviors, psychosocial interven-
er et al., 2009). Many of the same immunosenescence-related man- tions may also reduce the rise in cortisol and attenuate the decline
ifestations, like decreased numbers of naïve T cells, increased in DHEA, thus decreasing inflammatory effects. All these changes

Downloaded for dr laqif abdurrahman (laqif.abdurrahman@gmail.com) at Universitas Indonesia from ClinicalKey.com by Elsevier on May 20, 2018.
For personal use only. No other uses without permission. Copyright ©2018. Elsevier Inc. All rights reserved.
L. Müller, G. Pawelec / Brain, Behavior, and Immunity 39 (2014) 8–22 19

may possibly result in better immune responses (Jeckel et al., whether the benefits of this intervention are extendable to all pop-
2010). For example, following a psychological enrichment pro- ulations including frail, older people (Kohut and Senchina, 2004;
gram, which was designed to counteract social isolation and pas- Hong, 2011; Simpson et al., 2012).
sivity by increasing social activation, competence, and Taken together, the data obtained from different studies and
independence, elderly individuals showed increased levels of interventions suggest that both immune and neuroendocrine sys-
DHEA, testosterone, estradiol and growth hormone (Arnetz et al., tems are plastic and immunosenescence can be attenuated
1983). through different kinds of behavioral, psychological and nutritional
Another study demonstrated that older adults who practiced interventions.
relaxation techniques had reduced antibody titers to latent herpes
simplex virus-1 (HSV-1). Lower viral antibody titers suggest fewer
episodes of viral-reactivation, implying that this lifestyle interven- 10. Conclusions
tion would be associated with reduced chronic antigenic stimula-
tion (Gouin et al., 2008). Another study suggests that stress- A current evolutionary understanding of immunosenescence as
buffering strategies can lead to an improvement in cellular immu- postulated in this review is based on the consensus that the con-
nity involved in the control of latent viruses (Bauer et al., 2013). stantly remodelling immune system is most dynamic in childhood
Individuals with a positive type of personally have been found to as a result of responding to and generating memory for the diver-
have lower plasma cortisol levels and lower concentrations of IL- sity of pathogens to which the person is most exposed in early life
6 und CRP. These reduced levels of inflammatory markers have (in order to survive childhood and still be around in later life). The
been favourably associated with blood pressure and heart rate individual therefore invests heavily in expensive (and potentially
(Grant et al., 2009). Also, the outcome of the immune response to dangerous) adaptive immune responses to the most prevalent
vaccination (even in stressed elderly) appeared to be influenced pathogens. By puberty, the individuaĺs supply of naïve lympho-
by better social support (Glaser et al., 1992). Therefore, stress man- cytes will have been converted into memory cells for the most
agement interventions and resilience factors seem to modulate dif- important pathogens in that particular environment, increasing
ferent aspects of immunosenescence, most likely by modulation of protection and fitness for reproduction. Further investments in
neuroendocrine factors. generating more naïve cells, especially in the resource-intensive
It seems that regular exercise may play an important role also process of T-cell production, are no longer worthwhile and are also
for stress-reducing purposes. Physical activity is associated with dangerous in terms of immune pathology and autoimmunity. This
many advantageous effects on stress management, immunity, is the reason why output of lymphocytes from the bone marrow is
well-being and health. Regular physical activity attenuates neural reduced (but myeloid cells maintained or increased to facilitate
responses to stress in brain circuits, which are responsible for reg- sufficient innate immunity) and thymic involution occurs at pub-
ulating peripheral sympathetic activity, and could contribute to erty to greatly reduce naïve T-cell processing). In the modern
reductions in clinical disorders such as hypertension, heart failure, world, however, most people commonly survive for much longer
oxidative stress, and suppression of immunity (Dishman et al., than in the past and may be exposed to emerging pathogens, or be-
2006). It also has the important advantage of being non-invasive, come exposed to new pathogens by travel. The immune system is
low-cost and easy to implement (Bauer et al., 2013). For example, simply not evolved to cope with these challenges. It is less remark-
elderly individuals committed to a physical activity program dem- able that immunosenescence occurs than it is that there is so much
onstrated low levels of emotional distress (Wrosch et al., 2002) and reserve capacity in our immune systems that function may still be
had more moderate concentrations of salivary cortisol (Wrosch retained for over 100 years. The impact of behavioral variations on
et al., 2007). Higher levels of blood NK cell activity were character- immune status within this ‘‘narrow window of opportunity’’ may
istic for individuals with a relaxed life style and healthy mental nonetheless be great but by the nature of things cannot overcome
condition. Physically active behaviors would, thus, improve the the intrinsic programming which evolved during speciation. A
quantity, as well as the quality of blood NK cells, and, in turn, im- prime example of how we should not generalize too much comes
prove the mental health status by restraining anxiety and avoiding from our own work on exceptional survivors in the Leiden
depression (Boscolo et al., 2008). 85 + Study. Here, contrary to the generalized statement that immu-
Moderate-intensity exercise has been reported to be associated nosenescence is characterized by few naïve T cells, more memory
with anti-inflammatory effects, including lowering serum TNF and cells and higher pro-inflammatory status and that this is associated
IL-6 levels, and improving IL-10 and Treg counts (Simpson et al., with worse survival, we found the opposite: under the circum-
2008). Active lifestyle and regular exercise appear to provide pro- stances in which these people were living, having fewer naïve T
tection against dementia and cognitive decline, increasing the re- cells and more memory cells (responding to CMV antigens in a
lease of growth hormone, epinephrine, cortisol, prolactin, and pro-inflammatory manner) was prospectively associated with a
other factors, which have immunomodulatory effects (Handschin significant 8-year survival benefit (Derhovanessian et al., 2013).
and Spiegelman, 2008; Pedersen, 2011). Moreover, moderate train- This underlines the importance of considering each study popula-
ing in the elderly has been shown to improve T cell proliferation, to tion in the context of the local situation. This ‘‘local situation’’ is
increase IL-2 production and expression of the IL-2 receptor on T obviously hard to define, but certainly includes nutritional factors,
cells, to reduce the frequency of senescent T cells, and has been pathogen exposure, stresses of multiple types, exposures to envi-
associated with longer leukocyte telomeres and better in vivo im- ronmental pollutants, amount and type of exercise and other ex-
mune responses to vaccines (Simpson and Guy, 2010). In addition, tremely heterogeneous factors that will be very difficult or
physical activity can improve mucosal immune responses in the el- impossible to control for in people. It is our conviction that further
derly and has been shown to increase resistance to upper respira- insight into which of these multifarious factors may be most
tory infections (Sakamoto et al., 2009). Thus, the enhanced mucosal important for assuring maintenance of appropriate immunity into
immune response together with improved T-cell responses may late life will only be obtained by establishing large databases
build up stronger immunity to control bacterial and viral infec- attempting to correlate as many of these factors as possible in hu-
tions. However, many questions remain to be investigated, includ- man populations, and that animal models are unlikely to be partic-
ing the mechanisms which are involved in these processes, the ularly helpful in this respect. For example, to this end, the Berlin
appropriate type and dose of exercise, the potential clinical impact BASE II study is surveying 2200 Berlin residents at early middle
in dependency to this type and this particular dose, and finally, age and older age and collating data on psychological, cognitive,

Downloaded for dr laqif abdurrahman (laqif.abdurrahman@gmail.com) at Universitas Indonesia from ClinicalKey.com by Elsevier on May 20, 2018.
For personal use only. No other uses without permission. Copyright ©2018. Elsevier Inc. All rights reserved.
20 L. Müller, G. Pawelec / Brain, Behavior, and Immunity 39 (2014) 8–22

medical, genetic, nutritional, psychosocial and immunological fac- Boge, T., Remigy, M., Vaudaine, S., Tanguy, J., Bourdet-Sicard, R., van der Werf, S.,
2009. A probiotic fermented dairy drink improves antibody response to
tors in an effort to establish baseline conditions informative for
influenza vaccination in the elderly in two randomised controlled trials.
health and longevity over the lifespan (Bertram et al., 2013). Vaccine 27, 5677–5684.
Bonilla, F.A., Oettgen, H.C., 2010. Adaptive immunity. J. Allergy Clin. Immunol. 125,
S33–S40.
Acknowledgments Boscolo, P., Youinou, P., Theoharides, T.C., Cerulli, G., Conti, P., 2008. Environmental
and occupational stress and autoimmunity. Autoimmun. Rev. 7, 340–343.
This work was supported by the German Federal Ministry of Butcher, S.K., Killampalli, V., Lascelles, D., Wang, K., Alpar, E.K., Lord, J.M., 2005.
Raised cortisol:DHEAS ratios in the elderly after injury: potential impact upon
Education and Research (Bundesministerium für Bildung und Fors- neutrophil function and immunity. Aging Cell 4, 319–324.
chung, BMBF) under grant numbers #16SV5536K, #16SV5537, Calvanese, V., Lara, E., Kahn, A., Fraga, M.F., 2009. The role of epigenetics in aging
#16SV5538, and #16SV5837 and the BMBF Network ‘‘Geronto- and age-related diseases. Ageing Res. Rev. 8, 268–276.
Campbell, J.P., Riddell, N.E., Burns, V.E., Turner, M., van Zanten, J.J., Drayson, M.T.,
Shield’’ ((BMBF Gerontoshield 0315890F), as well as the European
Bosch, J.A., 2009. Acute exercise mobilises CD8+ T lymphocytes exhibiting an
Commission (EUFP7 IDEAL 259679) (to G.P.). LM would like to effector-memory phenotype. Brain Behav. Immun. 23, 767–775.
gratefully aknowlege Dr. Frank Schmidt for his competent help Chadalapaka, G., Jutooru, I., McAlees, A., Stefanac, T., Safe, S., 2008. Structure-
during the writing of the manuscript. dependent inhibition of bladder and pancreatic cancer cell growth by 2-
substituted glycyrrhetinic and ursolic acid derivatives. Bioorg. Med. Chem. Lett.
18, 2633–2639.
References Chainy, G.B., Manna, S.K., Chaturvedi, M.M., Aggarwal, B.B., 2000. Anethole blocks
both early and late cellular responses transduced by tumor necrosis factor:
effect on NF-kappaB, AP-1, JNK, MAPKK and apoptosis. Oncogene 19, 2943–
Adema, G.J., 2009. Dendritic cells from bench to bedside and back. Immunol. Lett.
2950.
122, 128–130.
Chang, S.S., Weiss, C.O., Xue, Q.L., Fried, L.P., 2012. Association between
Ademokun, A., Wu, Y.C., Dunn-Walters, D., 2010. The ageing B cell population:
inflammatory-related disease burden and frailty: results from the Women’s
composition and function. Biogerontology 11, 125–137.
Health and Aging Studies (WHAS) I and II. Arch. Gerontol. Geriatr. 54, 9–15.
Adzersen, K.H., Jess, P., Freivogel, K.W., Gerhard, I., Bastert, G., 2003. Raw and cooked
Chasapis, C.T., Loutsidou, A.C., Spiliopoulou, C.A., Stefanidou, M.E., 2012. Zinc and
vegetables, fruits, selected micronutrients, and breast cancer risk: a case-
human health: an update. Arch. Toxicol 86, 521–534.
control study in Germany. Nutr. Cancer 46, 131–137.
Colonna-Romano, G., Bulati, M., Aquino, A., Vitello, S., Lio, D., Candore, G., Caruso, C.,
Aggarwal, B.B., Shishodia, S., 2006. Molecular targets of dietary agents for
2008. B cell immunosenescence in the elderly and in centenarians.
prevention and therapy of cancer. Biochem. Pharmacol. 71, 1397–1421.
Rejuvenation Res. 11, 433–439.
Ames, B.N., Wakimoto, P., 2002. Are vitamin and mineral deficiencies a major cancer
Compston, J.E., 2002. Bone marrow and bone: a functional unit. J. Endocrinol. 173,
risk? Nat. Rev. Cancer 2, 694–704.
387–394.
Anderson, R.M., Weindruch, R., 2012. The caloric restriction paradigm: implications
Cooper, M.D., 2010. 99th Dahlem conference on infection, inflammation and chronic
for healthy human aging. Am. J. Hum. Biol. 24, 101–106.
inflammatory disorders: evolution of adaptive immunity in vertebrates. Clin.
Andersson, D., Liu, J.J., Nilsson, A., Duan, R.D., 2003. Ursolic acid inhibits
Exp. Immunol. 160, 58–61.
proliferation and stimulates apoptosis in HT29 cells following activation of
Corona, A.W., Fenn, A.M., Godbout, J.P., 2012. Cognitive and behavioral
alkaline sphingomyelinase. Anticancer Res. 23, 3317–3322.
consequences of impaired immunoregulation in aging. J. Neuroimmun.
Appay, V., Sauce, D., Prelog, M., 2010. The role of the thymus in immunosenescence.
Pharmacol. 7, 7–23.
lessons from the study of thymectomized individuals. Aging (Albany NY) 2, 78–
Damjanovic, A.K., Yang, Y., Glaser, R., Kiecolt-Glaser, J.K., Nguyen, H., Laskowski, B.,
81.
Zou, Y., Beversdorf, D.Q., Weng, N.P., 2007. Accelerated telomere erosion is
Arnetz, B.B., Theorell, T., Levi, L., Kallner, A., Eneroth, P., 1983. An experimental
associated with a declining immune function of caregivers of Alzheimer’s
study of social isolation of elderly people: psychoendocrine and metabolic
disease patients. J. Immunol. 179, 4249–4254.
effects. Psychosom. Med. 45, 395–406.
Deeks, S.G., Verdin, E., McCune, J.M., 2012. Immunosenescence and HIV. Curr. Opin.
Arnold, C.R., Wolf, J., Brunner, S., Herndler-Brandstetter, D., Grubeck-Loebenstein, B.,
Immunol. 24, 501–506.
2011. Gain and loss of T cell subsets in old age–age-related reshaping of the T
Delarosa, O., Tarazona, R., Casado, J.G., Alonso, C., Ostos, B., Pena, J., Solana, R., 2002.
cell repertoire. J. Clin. Immunol. 31, 137–146.
Valpha24+ NKT cells are decreased in elderly humans. Exp. Gerontol. 37, 213–
Ashwell, J.D., Lu, F.W., Vacchio, M.S., 2000. Glucocorticoids in T cell development
217.
and function⁄. Annu. Rev. Immunol. 18, 309–345.
Della bella, S., Bierti, L., Presicce, P., Arienti, R., Valenti, M., Saresella, M., Vergani, C.,
Aspinall, R., Pitts, D., Lapenna, A., Mitchell, W., 2010. Immunity in the elderly: the
Villa, M.L., 2007. Peripheral blood dendritic cells and monocytes are differently
role of the thymus. J. Comp. Pathol. 142 (Suppl 1), S111–S115.
regulated in the elderly. Clin. Immunol. 122, 220–228.
Atarashi, K., Tanoue, T., Oshima, K., Suda, W., Nagano, Y., Nishikawa, H., Fukuda, S.,
Derhovanessian, E., Maier, A.B., Hahnel, K., Beck, R., de Craen, A.J., Slagboom, E.P.,
Saito, T., Narushima, S., Hase, K., Kim, S., Fritz, J.V., Wilmes, P., Ueha, S.,
Westendorp, R.G., Pawelec, G., 2011. Infection with cytomegalovirus but not
Matsushima, K., Ohno, H., Olle, B., Sakaguchi, S., Taniguchi, T., Morita, H.,
herpes simplex virus induces the accumulation of late-differentiated CD4+ and
Hattori, M., Honda, K., 2013. Treg induction by a rationally selected mixture of
CD8+ T-cells in humans. J. Gen. Virol. 92, 2746–2756.
Clostridia strains from the human microbiota. Nature 500, 232–236.
Derhovanessian, E., Solana, R., Larbi, A., Pawelec, G., 2008. Immunity, ageing and
Bartlett, D.B., Firth, C.M., Phillips, A.C., Moss, P., Baylis, D., Syddall, H., Sayer, A.A.,
cancer. Immun. Ageing 5, 11.
Cooper, C., Lord, J.M., 2012. The age-related increase in low-grade systemic
Derhovanessian, E., Theeten, H., Hahnel, K., van Damme, P., Cools, N., Pawelec, G.,
inflammation (Inflammaging) is not driven by cytomegalovirus infection. Aging
2013. Cytomegalovirus-associated accumulation of late-differentiated CD4 T-
Cell 11, 912–915.
cells correlates with poor humoral response to influenza vaccination. Vaccine
Bauer, M.E., 2005. Stress, glucocorticoids and ageing of the immune system. Stress
31, 685–690.
8, 69–83.
Dewan, S.K., Zheng, S.B., Xia, S.J., Bill, K., 2012. Senescent remodeling of the immune
Bauer, M.E., Jeckel, C.M., Luz, C., 2009. The role of stress factors during aging of the
system and its contribution to the predisposition of the elderly to infections.
immune system. Ann. N. Y. Acad. Sci. 1153, 139–152.
Chin. Med. J. (Engl.) 125, 3325–3331.
Bauer, M.E., Muller, G.C., Correa, B.L., Vianna, P., Turner, J.E., Bosch, J.A., 2013.
Dishman, R.K., Berthoud, H.R., Booth, F.W., Cotman, C.W., Edgerton, V.R., Fleshner,
Psychoneuroendocrine interventions aimed at attenuating immunosenescence.
M.R., Gandevia, S.C., Gomez-Pinilla, F., Greenwood, B.N., Hillman, C.H., Kramer,
a review. Biogerontology 14, 9–20.
A.F., Levin, B.E., Moran, T.H., Russo-Neustadt, A.A., Salamone, J.D., van
Beerman, I., Bhattacharya, D., Zandi, S., Sigvardsson, M., Weissman, I.L., Bryder, D.,
Hoomissen, J.D., Wade, C.E., York, D.A., Zigmond, M.J., 2006. Neurobiology of
Rossi, D.J., 2010. Functionally distinct hematopoietic stem cells modulate
exercise. Obesity (Silver Spring) 14, 345–356.
hematopoietic lineage potential during aging by a mechanism of clonal
Dock, J.N., Effros, R.B., 2011. Role of CD8 T Cell Replicative Senescence in Human
expansion. Proc. Natl. Acad. Sci. U.S.A. 107, 5465–5470.
Aging and in HIV-mediated Immunosenescence. Aging Dis. 2, 382–397.
Bennett, J.M., Glaser, R., Malarkey, W.B., Beversdorf, D.Q., Peng, J., Kiecolt-Glaser,
Dockray, S., Steptoe, A., 2010. Positive affect and psychobiological processes.
J.K., 2012. Inflammation and reactivation of latent herpesviruses in older adults.
Neurosci. Biobehav. Rev. 35, 69–75.
Brain Behav. Immun. 26, 739–746.
Duncan, S.H., Flint, H.J., 2013. Probiotics and prebiotics and health in ageing
Bertram, L., Bockenhoff, A., Demuth, I., Duzel, S., Eckardt, R., Li, S.C., Lindenberger, U.,
populations. Maturitas 75, 44–50.
Pawelec, G., Siedler, T., Wagner, G.G., Steinhagen-Thiessen, E., 2013. Cohort
Dykstra, B., de Haan, G., 2008. Hematopoietic stem cell aging and self-renewal. Cell
Profile: The Berlin Aging Study II (BASE-II). Int. J. Epidemiol. [Epub ahead of
Tissue Res. 331, 91–101.
print].
Effros, R.B., Dagarag, M., Spaulding, C., Man, J., 2005. The role of CD8+ T-cell
Biagi, E., Candela, M., Fairweather-Tait, S., Franceschi, C., Brigidi, P., 2012. Aging of
replicative senescence in human aging. Immunol. Rev. 205, 147–157.
the human metaorganism: the microbial counterpart. Age (Dordr) 34, 247–267.
Elenkov, I.J., Chrousos, G.P., 1999. Stress hormones, Th1/Th2 patterns, pro/anti-
Biagi, E., Nylund, L., Candela, M., Ostan, R., Bucci, L., Pini, E., Nikkila, J., Monti, D.,
inflammatory cytokines and susceptibility to disease. Trends Endocrinol. Metab.
Satokari, R., Franceschi, C., Brigidi, P., de Vos, W., 2010. Through ageing, and
10, 359–368.
beyond: gut microbiota and inflammatory status in seniors and centenarians.
Es-Saady, D., Simon, A., Ollier, M., Maurizis, J.C., Chulia, A.J., Delage, C., 1996.
PLoS One 5, e10667.
Inhibitory effect of ursolic acid on B16 proliferation through cell cycle arrest.
Blum, S., Haller, D., Pfeifer, A., Schiffrin, E.J., 2002. Probiotics and immune response.
Cancer Lett. 106, 193–197.
Clin. Rev. Allergy Immunol. 22, 287–309.

Downloaded for dr laqif abdurrahman (laqif.abdurrahman@gmail.com) at Universitas Indonesia from ClinicalKey.com by Elsevier on May 20, 2018.
For personal use only. No other uses without permission. Copyright ©2018. Elsevier Inc. All rights reserved.
L. Müller, G. Pawelec / Brain, Behavior, and Immunity 39 (2014) 8–22 21

Fernandez-Morera, J.L., Calvanese, V., Rodriguez-Rodero, S., Menendez-Torre, E., Kiecolt-Glaser, J.K., Preacher, K.J., Maccallum, R.C., Atkinson, C., Malarkey, W.B.,
Fraga, M.F., 2010. Epigenetic regulation of the immune system in health and Glaser, R., 2003. Chronic stress and age-related increases in the
disease. Tissue Antigens 76, 431–439. proinflammatory cytokine IL-6. Proc. Natl. Acad. Sci. U.S.A. 100, 9090–9095.
Ferrando-Martinez, S., Ruiz-Mateos, E., Hernandez, A., Gutierrez, E., Rodriguez- Kim, D.H., Kim, C.H., Kim, M.S., Kim, J.Y., Jung, K.J., Chung, J.H., An, W.G., Lee, J.W., Yu,
Mendez Mdel, M., Ordonez, A., leal, M., 2011. Age-related deregulation of naive B.P., Chung, H.Y., 2007. Suppression of age-related inflammatory NF-kappaB
T cell homeostasis in elderly humans. Age (Dordr) 33, 197–207. activation by cinnamaldehyde. Biogerontology 8, 545–554.
Ford, E.S., Bergmann, M.M., Kroger, J., Schienkiewitz, A., Weikert, C., Boeing, H., Kim, S.S., Oh, O.J., Min, H.Y., Park, E.J., Kim, Y., Park, H.J., Nam Han, Y., Lee, S.K., 2003.
2009. Healthy living is the best revenge: findings from the European Eugenol suppresses cyclooxygenase-2 expression in lipopolysaccharide-
Prospective Investigation Into Cancer and Nutrition-Potsdam study. Arch. stimulated mouse macrophage RAW264.7 cells. Life Sci. 73, 337–348.
Intern. Med. 169, 1355–1362. Kohler, S., Wagner, U., Pierer, M., Kimmig, S., Oppmann, B., Mowes, B., Julke, K.,
Franceschi, C., Bonafe, M., Valensin, S., Olivieri, F., de Luca, M., Ottaviani, E., de Romagnani, C., Thiel, A., 2005. Post-thymic in vivo proliferation of naive CD4+ T
Benedictis, G., 2000. Inflamm-aging. An evolutionary perspective on cells constrains the TCR repertoire in healthy human adults. Eur. J. Immunol. 35,
immunosenescence. Ann. N. Y. Acad. Sci. 908, 244–254. 1987–1994.
Franceschi, C., Capri, M., Monti, D., Giunta, S., Olivieri, F., Sevini, F., Panourgia, M.P., Kohut, M.L., Cooper, M.M., Nickolaus, M.S., Russell, D.R., Cunnick, J.E., 2002. Exercise
Invidia, L., Celani, L., Scurti, M., Cevenini, E., Castellani, G.C., Salvioli, S., 2007. and psychosocial factors modulate immunity to influenza vaccine in elderly
Inflammaging and anti-inflammaging: a systemic perspective on aging individuals. J. Gerontol. A Biol. Sci. Med. Sci. 57, M557–M562.
and longevity emerged from studies in humans. Mech. Ageing Dev. 128, 92– Kohut, M.L., Senchina, D.S., 2004. Reversing age-associated immunosenescence via
105. exercise. Exerc. Immunol. Rev. 10, 6–41.
Fuente Mde, L., Cruces, J., Hernandez, O., Ortega, E., 2011. Strategies to improve the Kruger, K., Frost, S., Most, E., Volker, K., Pallauf, J., Mooren, F.C., 2009. Exercise affects
functions and redox state of the immune system in aged subjects. Curr. Pharm. tissue lymphocyte apoptosis via redox-sensitive and Fas-dependent signaling
Des. 17, 3966–3993. pathways. Am. J. Physiol. Regul. Integr. Comp. Physiol. 296, R1518–R1527.
Fulop, T., Larbi, A., Douziech, N., Fortin, C., Guerard, K.P., Lesur, O., Khalil, A., Dupuis, Kwak, J.H., Lee, J.H., Ahn, C.W., Park, S.H., Shim, S.T., Song, Y.D., Han, E.N., Lee, K.H.,
G., 2004. Signal transduction and functional changes in neutrophils with aging. Chae, J.S., 2010. Black soy peptide supplementation improves glucose control in
Aging Cell 3, 217–226. subjects with prediabetes and newly diagnosed type 2 diabetes mellitus. J. Med.
Geiger, H., de Haan, G., Florian, M.C., 2013. The ageing haematopoietic stem cell Food 13, 1307–1312.
compartment. Nat. Rev. Immunol. 13, 376–389. Lang, P.O., Govind, S., Aspinall, R., 2013. Reversing T cell immunosenescence. why,
Glaser, R., Kiecolt-Glaser, J.K., Bonneau, R.H., Malarkey, W., Kennedy, S., Hughes, J., who, and how. Age (Dordr) 35, 609–620.
1992. Stress-induced modulation of the immune response to recombinant Lanier, L.L., Sun, J.C., 2009. Do the terms innate and adaptive immunity create
hepatitis B vaccine. Psychosom. Med. 54, 22–29. conceptual barriers? Nat. Rev. Immunol. 9, 302–303.
Gleeson, M., Bishop, N.C., Stensel, D.J., Lindley, M.R., Mastana, S.S., Nimmo, M.A., Larbi, A., Franceschi, C., Mazzatti, D., Solana, R., Wikby, A., Pawelec, G., 2008. Aging
2011. The anti-inflammatory effects of exercise: mechanisms and implications of the immune system as a prognostic factor for human longevity. Physiology
for the prevention and treatment of disease. Nat. Rev. Immunol. 11, 607–615. (Bethesda) 23, 64–74.
Globerson, A., Effros, R.B., 2000. Ageing of lymphocytes and lymphocytes in the Larbi, A., Pawelec, G., Wong, S.C., Goldeck, D., Tai, J.J., Fulop, T., 2011. Impact of age
aged. Immunol. Today 21, 515–521. on T cell signaling: a general defect or specific alterations? Ageing Res. Rev. 10,
Gonzalo, S., 2010. Epigenetic alterations in aging. J. Appl. Physiol. 109, 586–597. 370–378.
Goodwin, K., Viboud, C., Simonsen, L., 2006. Antibody response to influenza Li, Y., Daniel, M., Tollefsbol, T.O., 2011. Epigenetic regulation of caloric restriction in
vaccination in the elderly: a quantitative review. Vaccine 24, 1159–1169. aging. BMC. Med. 9, 98.
Gouin, J.P., Hantsoo, L., Kiecolt-Glaser, J.K., 2008. Immune dysregulation and chronic Liu, K., Catalfamo, M., Li, Y., Henkart, P.A., Weng, N.P., 2002. IL-15 mimics T cell
stress among older adults: a review. Neuroimmunomodulation 15, 251–259. receptor crosslinking in the induction of cellular proliferation, gene expression,
Grant, N., Hamer, M., Steptoe, A., 2009. Social isolation and stress-related and cytotoxicity in CD8+ memory T cells. Proc. Natl. Acad. Sci. U.S.A. 99, 6192–
cardiovascular, lipid, and cortisol responses. Ann. Behav. Med. 37, 29–37. 6197.
Gruver, A.L., Hudson, L.L., Sempowski, G.D., 2007. Immunosenescence of ageing. J. Luz, C., Dornelles, F., Preissler, T., Collaziol, D., da Cruz, I.M., Bauer, M.E., 2003.
Pathol. 211, 144–156. Impact of psychological and endocrine factors on cytokine production of
Gui, J., Mustachio, L.M., Su, D.M., Craig, R.W., 2012. Thymus size and age-related healthy elderly people. Mech. Ageing Dev. 124, 887–895.
thymic involution: early programming, sexual dimorphism, progenitors and Lynch, H.E., Goldberg, G.L., Chidgey, A., van den Brink, M.R., Boyd, R., Sempowski,
stroma. Aging Dis. 3, 280–290. G.D., 2009. Thymic involution and immune reconstitution. Trends Immunol. 30,
Guigoz, Y., Dore, J., Schiffrin, E.J., 2008. The inflammatory status of old age can be 366–373.
nurtured from the intestinal environment. Curr. Opin. Clin. Nutr. Metab. Care Mahbub, S., Brubaker, A.L., Kovacs, E.J., 2011. Aging of the innate immune system:
11, 13–20. an update. Curr. Immunol. Rev. 7, 104–115.
Hajat, A., Diez-Roux, A., Franklin, T.G., Seeman, T., Shrager, S., Ranjit, N., Castro, C., Maninger, N., Wolkowitz, O.M., Reus, V.I., Epel, E.S., Mellon, S.H., 2009.
Watson, K., Sanchez, B., Kirschbaum, C., 2010. Socioeconomic and race/ethnic Neurobiological and neuropsychiatric effects of dehydroepiandrosterone
differences in daily salivary cortisol profiles: the multi-ethnic study of (DHEA) and DHEA sulfate (DHEAS). Front. Neuroendocrinol. 30, 65–91.
atherosclerosis. Psychoneuroendocrinology 35, 932–943. Medzhitov, R., Janeway Jr., C.A., 1997. Innate immunity: the virtues of a nonclonal
Handschin, C., Spiegelman, B.M., 2008. The role of exercise and PGC1alpha in system of recognition. Cell 91, 295–298.
inflammation and chronic disease. Nature 454, 463–469. Meydani, S.N., Barklund, M.P., Liu, S., Meydani, M., Miller, R.A., Cannon, J.G., Morrow,
Harmand, P.O., Duval, R., Delage, C., Simon, A., 2005. Ursolic acid induces apoptosis F.D., Rocklin, R., Blumberg, J.B., 1990. Vitamin E supplementation enhances cell-
through mitochondrial intrinsic pathway and caspase-3 activation in M4Beu mediated immunity in healthy elderly subjects. Am. J. Clin. Nutr. 52, 557–563.
melanoma cells. Int. J. Cancer 114, 1–11. Meydani, S.N., Ha, W.K., 2000. Immunologic effects of yogurt. Am. J. Clin. Nutr. 71,
Haugen, F., Norheim, F., Lian, H., Wensaas, A.J., Dueland, S., Berg, O., Funderud, A., 861–872.
Skalhegg, B.S., Raastad, T., Drevon, C.A., 2010. IL-7 is expressed and secreted by Mocchegiani, E., Muzzioli, M., Giacconi, R., Cipriano, C., Gasparini, N., Franceschi, C.,
human skeletal muscle cells. Am J Physiol Cell Physiol 298, C807–C816. Gaetti, R., Cavalieri, E., Suzuki, H., 2003. Metallothioneins/PARP-1/IL-6 interplay
Hawkley, L.C., Cacioppo, J.T., 2004. Stress and the aging immune system. Brain on natural killer cell activity in elderly: parallelism with nonagenarians and old
Behav. Immun. 18, 114–119. infected humans. Effect of zinc supply. Mech. Ageing Dev. 124, 459–468.
Hazeldine, J., Hampson, P., Lord, J.M., 2012. Reduced release and binding of perforin Molano, A., Meydani, S.N., 2012. Vitamin E, signalosomes and gene expression in T
at the immunological synapse underlies the age-related decline in natural killer cells. Mol. Aspects Med. 33, 55–62.
cell cytotoxicity. Aging Cell 11, 751–759. Morimoto, S.S., Alexopoulos, G.S., 2011. Immunity, aging, and geriatric depression.
Heffner, K.L., 2011. Neuroendocrine effects of stress on immunity in the elderly: Psychiatr. Clin. North Am. 34, 437–449, ix.
implications for inflammatory disease. Immunol. Allergy Clin. North Am. 31, Müller, L., Fulop, T., Pawelec, G., 2013. Immunosenescence in vertebrates and
95–108. invertebrates. Immun. Ageing 10, 12.
Hong, S., 2011. Can we jog our way to a younger-looking immune system? Brain Muzzioli, M., Stecconi, R., Moresi, R., Provinciali, M., 2009. Zinc improves the
Behav. Immun. 25, 1519–1520. development of human CD34+ cell progenitors towards NK cells and increases
Hsu, Y.L., Kuo, P.L., Lin, C.C., 2004. Proliferative inhibition, cell-cycle dysregulation, the expression of GATA-3 transcription factor in young and old ages.
and induction of apoptosis by ursolic acid in human non-small cell lung cancer Biogerontology 10, 593–604.
A549 cells. Life Sci. 75, 2303–2316. Nazmi, A., Diez Roux, A., Ranjit, N., Seeman, T.E., Jenny, N.S., 2010. Cross-sectional
Jeckel, C.M., Lopes, R.P., Berleze, M.C., Luz, C., Feix, L., Argimon, I.I., Stein, L.M., Bauer, and longitudinal associations of neighborhood characteristics with
M.E., 2010. Neuroendocrine and immunological correlates of chronic stress in inflammatory markers: findings from the multi-ethnic study of
‘strictly healthy’ populations. Neuroimmunomodulation 17, 9–18. atherosclerosis. Health Place 16, 1104–1112.
Jenny, N.S., 2012. Inflammation in aging: cause, effect, or both? Discov. Med. 13, Nikolich-Zugich, J., Messaoudi, I., 2005. Mice and flies and monkeys too: caloric
451–460. restriction rejuvenates the aging immune system of non-human primates. Exp.
Jolly, C.A., 2007. Is dietary restriction beneficial for human health, such as for Gerontol. 40, 884–893.
immune function? Curr. Opin. Lipidol. 18, 53–57. Nussinovitch, U., Shoenfeld, Y., 2012. The role of gender and organ specific
Keller, R., 1993. The macrophage response to infectious agents: mechanisms of autoimmunity. Autoimmun. Rev. 11, A377–A385.
macrophage activation and tumour cell killing. Res. Immunol. 144, 271–273, Ornish, D., Lin, J., Chan, J.M., Epel, E., Kemp, C., Weidner, G., Marlin, R., Frenda, S.J.,
discussion 294-8. Magbanua, M.J., Daubenmier, J., Estay, I., Hills, N.K., Chainani-Wu, N., Carroll,
Khanfer, R., Lord, J.M., Phillips, A.C., 2011. Neutrophil function and cortisol:DHEAS P.R., Blackburn, E.H., 2013. Effect of comprehensive lifestyle changes on
ratio in bereaved older adults. Brain Behav. Immun. 25, 1182–1186. telomerase activity and telomere length in men with biopsy-proven low-risk

Downloaded for dr laqif abdurrahman (laqif.abdurrahman@gmail.com) at Universitas Indonesia from ClinicalKey.com by Elsevier on May 20, 2018.
For personal use only. No other uses without permission. Copyright ©2018. Elsevier Inc. All rights reserved.
22 L. Müller, G. Pawelec / Brain, Behavior, and Immunity 39 (2014) 8–22

prostate cancer: 5-year follow-up of a descriptive pilot study. Lancet Oncol. 14, Simpson, R.J., Lowder, T.W., Spielmann, G., Bigley, A.B., Lavoy, E.C., Kunz, H., 2012.
1112–1120. Exercise and the aging immune system. Ageing Res. Rev. 11, 404–420.
Ornish, D., Lin, J., Daubenmier, J., Weidner, G., Epel, E., Kemp, C., Magbanua, M.J., Solana, R., Pawelec, G., Tarazona, R., 2006. Aging and innate immunity. Immunity
Marlin, R., Yglecias, L., Carroll, P.R., Blackburn, E.H., 2008. Increased telomerase 24, 491–494.
activity and comprehensive lifestyle changes: a pilot study. Lancet Oncol. 9, Spielmann, G., McFarlin, B.K., O’Connor, D.P., Smith, P.J., Pircher, H., Simpson, R.J.,
1048–1057. 2011. Aerobic fitness is associated with lower proportions of senescent blood T-
Ouchi, N., Parker, J.L., Lugus, J.J., Walsh, K., 2011. Adipokines in inflammation and cells in man. Brain Behav. Immun. 25, 1521–1529.
metabolic disease. Nat. Rev. Immunol. 11, 85–97. Steensberg, A., Fischer, C.P., Keller, C., Moller, K., Pedersen, B.K., 2003. IL-6 enhances
Pae, M., Meydani, S.N., Wu, D., 2012. The role of nutrition in enhancing immunity in plasma IL-1ra, IL-10, and cortisol in humans. Am. J. Physiol. Endocrinol. Metab.
aging. Aging Dis. 3, 91–129. 285, E433–E437.
Pan, Z., Chang, C., 2012. Gender and the regulation of longevity: implications for Stewart, L.K., Flynn, M.G., Campbell, W.W., Craig, B.A., Robinson, J.P., McFarlin, B.K.,
autoimmunity. Autoimmun. Rev. 11, A393–A403. Timmerman, K.L., Coen, P.M., Felker, J., Talbert, E., 2005. Influence of exercise
Pathak, A.K., Bhutani, M., Nair, A.S., Ahn, K.S., Chakraborty, A., Kadara, H., Guha, S., training and age on CD14+ cell-surface expression of toll-like receptor 2 and 4.
Sethi, G., Aggarwal, B.B., 2007. Ursolic acid inhibits STAT3 activation pathway Brain Behav. Immun. 19, 389–397.
leading to suppression of proliferation and chemosensitization of human Takahashi, A., Hanson, M.G., Norell, H.R., Havelka, A.M., Kono, K., Malmberg, K.J.,
multiple myeloma cells. Mol. Cancer Res. 5, 943–955. Kiessling, R.V., 2005. Preferential cell death of CD8+ effector memory (CCR7-
Pawelec, G., 2012. Hallmarks of human ‘‘immunosenescence’’: adaptation or CD45RA-) T cells by hydrogen peroxide-induced oxidative stress. J. Immunol.
dysregulation? Immun. Ageing 9, 15. 174, 6080–6087.
Pawelec, G., Derhovanessian, E., Larbi, A., Strindhall, J., Wikby, A., 2009. Takeda, K., Okumura, K., 2007. Effects of a fermented milk drink containing
Cytomegalovirus and human immunosenescence. Rev. Med. Virol. 19, 47–56. Lactobacillus casei strain Shirota on the human NK-cell activity. J. Nutr. 137,
Pedersen, B.K., 2011. Exercise-induced myokines and their role in chronic diseases. 791S–793S.
Brain Behav. Immun. 25, 811–816. Timmerman, K.L., Flynn, M.G., Coen, P.M., Markofski, M.M., Pence, B.D., 2008.
Peralbo, E., Delarosa, O., Gayoso, I., Pita, M.L., Tarazona, R., Solana, R., 2006. Exercise training-induced lowering of inflammatory (CD14+CD16+) monocytes:
Decreased frequency and proliferative response of invariant Valpha24Vbeta11 a role in the anti-inflammatory influence of exercise? J. Leukoc. Biol. 84, 1271–
natural killer T (iNKT) cells in healthy elderly. Biogerontology 7, 483–492. 1278.
Prasad, S., Sung, B., Aggarwal, B.B., 2012. Age-associated chronic diseases require Tower, J., Arbeitman, M., 2009. The genetics of gender and life span. J. Biol. 8, 38.
age-old medicine: role of chronic inflammation. Prev. Med. 54 (Suppl), S29–S37. Tripathi, S., Maier, K.G., Bruch, D., Kittur, D.S., 2007. Effect of 6-gingerol on pro-
Rickman, A.D., Williamson, D.A., Martin, C.K., Gilhooly, C.H., Stein, R.I., Bales, C.W., inflammatory cytokine production and costimulatory molecule expression in
Roberts, S., Das, S.K., 2011. The CALERIE Study: design and methods of an murine peritoneal macrophages. J. Surg. Res. 138, 209–213.
innovative 25% caloric restriction intervention. Contemp. Clin. Trials 32, 874–881. Trzonkowski, P., Szmit, E., Mysliwska, J., Mysliwski, A., 2006. CD4+CD25+ T
Rymkiewicz, P.D., Heng, Y.X., Vasudev, A., Larbi, A., 2012. The immune system in the regulatory cells inhibit cytotoxic activity of CTL and NK cells in humans-
aging human. Immunol. Res. 53, 235–250. impact of immunosenescence. Clin. Immunol. 119, 307–316.
Sakamoto, Y., Ueki, S., Kasai, T., Takato, J., Shimanuki, H., Honda, H., Ito, T., Haga, H., Turner, J.E., Aldred, S., Witard, O.C., Drayson, M.T., Moss, P.M., Bosch, J.A., 2010.
2009. Effect of exercise, aging and functional capacity on acute secretory Latent cytomegalovirus infection amplifies CD8 T-lymphocyte and egress in
immunoglobulin: a response in elderly people over 75 years of age. Geriatr. response to exercise. Brain Behav. Immun. 24, 1362–1370.
Gerontol. Int. 9, 81–88. von Kanel, R., Dimsdale, J.E., Mills, P.J., Ancoli-Israel, S., Patterson, T.L., Mausbach,
Sauce, D., Appay, V., 2011. Altered thymic activity in early life: how does it affect the B.T., Grant, I., 2006. Effect of Alzheimer caregiving stress and age on frailty
immune system in young adults? Curr. Opin. Immunol. 23, 543–548. markers interleukin-6, C-reactive protein, and D-dimer. J. Gerontol. A Biol. Sci.
Schiffrin, E.J., Morley, J.E., Donnet-Hughes, A., Guigoz, Y., 2010. The inflammatory Med. Sci. 61, 963–969.
status of the elderly: the intestinal contribution. Mutat. Res. 690, 50–56. Vulevic, J., Drakoularakou, A., Yaqoob, P., Tzortzis, G., Gibson, G.R., 2008. Modulation
Schwarz, B.A., Bhandoola, A., 2006. Trafficking from the bone marrow to the of the fecal microflora profile and immune function by a novel trans-
thymus: a prerequisite for thymopoiesis. Immunol. Rev. 209, 47–57. galactooligosaccharide mixture (B-GOS) in healthy elderly volunteers. Am. J.
Sempowski, G.D., Hale, L.P., Sundy, J.S., Massey, J.M., Koup, R.A., Douek, D.C., Patel, Clin. Nutr. 88, 1438–1446.
D.D., Haynes, B.F., 2000. Leukemia inhibitory factor, oncostatin M, IL-6, and Wack, A., Cossarizza, A., Heltai, S., Barbieri, D., D’Addato, S., Fransceschi, C.,
stem cell factor mRNA expression in human thymus increases with age and is Dellabona, P., Casorati, G., 1998. Age-related modifications of the human
associated with thymic atrophy. J. Immunol. 164, 2180–2187. alphabeta T cell repertoire due to different clonal expansions in the CD4+ and
Shanley, D.P., Aw, D., Manley, N.R., Palmer, D.B., 2009. An evolutionary perspective CD8+ subsets. Int. Immunol. 10, 1281–1288.
on the mechanisms of immunosenescence. Trends Immunol. 30, 374–381. Walsh, N.P., Gleeson, M., Shephard, R.J., Gleeson, M., Woods, J.A., Bishop, N.C.,
Shaw, A.C., Joshi, S., Greenwood, H., Panda, A., Lord, J.M., 2010. Aging of the innate Fleshner, M., Green, C., Pedersen, B.K., Hoffman-Goetz, L., Rogers, C.J., Northoff,
immune system. Curr. Opin. Immunol. 22, 507–513. H., Abbasi, A., Simon, P., 2011. Position statement. Part one: immune function
Sheikh, A., Shamsuzzaman, S., Ahmad, S.M., Nasrin, D., Nahar, S., Alam, M.M., Al and exercise. Exerc. Immunol. Rev. 17, 6–63.
tarique, A., Begum, Y.A., Qadri, S.S., Chowdhury, M.I., Saha, A., Larson, C.P., Qadri, Wang, J.S., Lin, C.T., 2010. Systemic hypoxia promotes lymphocyte apoptosis
F., 2010. Zinc influences innate immune responses in children with induced by oxidative stress during moderate exercise. Eur. J. Appl. Physiol. 108,
enterotoxigenic Escherichia coli-induced diarrhea. J. Nutr. 140, 1049–1056. 371–382.
Shiels, P.G., McGlynn, L.M., Macintyre, A., Johnson, P.C., Batty, G.D., Burns, H., Warren, L.A., Rossi, D.J., 2009. Stem cells and aging in the hematopoietic system.
Cavanagh, J., Deans, K.A., Ford, I., McConnachie, A., McGinty, A., McLean, J.S., Mech. Ageing Dev. 130, 46–53.
Millar, K., Sattar, N., Tannahill, C., Velupillai, Y.N., Packard, C.J., 2011. Accelerated Weiskopf, D., Weinberger, B., Grubeck-Loebenstein, B., 2009. The aging of the
telomere attrition is associated with relative household income, diet and immune system. Transpl. Int. 22, 1041–1050.
inflammation in the pSoBid cohort. PLoS One 6, e22521. Wellinghausen, N., Martin, M., Rink, L., 1997. Zinc inhibits interleukin-1-dependent
Shimizu, K., Kimura, F., Akimoto, T., Akama, T., Tanabe, K., Nishijima, T., Kuno, S., T cell stimulation. Eur. J. Immunol. 27, 2529–2535.
Kono, I., 2008. Effect of moderate exercise training on T-helper cell Woods, J.A., Keylock, K.T., Lowder, T., Vieira, V.J., Zelkovich, W., Dumich, S.,
subpopulations in elderly people. Exerc. Immunol. Rev. 14, 24–37. Colantuano, K., Lyons, K., Leifheit, K., Cook, M., Chapman-Novakofski, K.,
Shimizu, K., Suzuki, N., Imai, T., Aizawa, K., Nanba, H., Hanaoka, Y., Kuno, S., Mesaki, McAuley, E., 2009. Cardiovascular exercise training extends influenza vaccine
N., Kono, I., Akama, T., 2011. Monocyte and T-cell responses to exercise training seroprotection in sedentary older adults: the immune function intervention
in elderly subjects. J. Strength Cond. Res. 25, 2565–2572. trial. J. Am. Geriatr. Soc. 57, 2183–2191.
Shishodia, S., Potdar, P., Gairola, C.G., Aggarwal, B.B., 2003. Curcumin Wrosch, C., Schulz, R., Heckhausen, J., 2002. Health stresses and depressive
(diferuloylmethane) down-regulates cigarette smoke-induced NF-kappaB symptomatology in the elderly: the importance of health engagement control
activation through inhibition of IkappaBalpha kinase in human lung epithelial strategies. Health Psychol. 21, 340–348.
cells: correlation with suppression of COX-2, MMP-9 and cyclin D1. Wrosch, C., Schulz, R., Miller, G.E., Lupien, S., Dunne, E., 2007. Physical health
Carcinogenesis 24, 1269–1279. problems, depressive mood, and cortisol secretion in old age: buffer effects of
Simpson, R.J., 2011. Aging, persistent viral infections, and immunosenescence. Can health engagement control strategies. Health Psychol. 26, 341–349.
exercise ‘‘make space’’? Exerc. Sport Sci. Rev. 39, 23–33. Yeh, S.H., Chuang, H., Lin, L.W., Hsiao, C.Y., Eng, H.L., 2006. Regular tai chi chuan
Simpson, R.J., Cosgrove, C., Ingram, L.A., Florida-James, G.D., Whyte, G.P., Pircher, H., exercise enhances functional mobility and CD4CD25 regulatory T cells. Br. J.
Guy, K., 2008. Senescent T-lymphocytes are mobilised into the peripheral blood Sports Med. 40, 239–243.
compartment in young and older humans after exhaustive exercise. Brain Yim, E.K., Lee, K.H., Namkoong, S.E., Um, S.J., Park, J.S., 2006. Proteomic analysis of
Behav. Immun. 22, 544–551. ursolic acid-induced apoptosis in cervical carcinoma cells. Cancer Lett. 235,
Simpson, R.J., Florida-James, G.D., Cosgrove, C., Whyte, G.P., Macrae, S., Pircher, H., 209–220.
Guy, K., 2007. High-intensity exercise elicits the mobilization of senescent T Zhang, Y.X., Kong, C.Z., Wang, L.H., Li, J.Y., Liu, X.K., Xu, B., Xu, C.L., Sun, Y.H., 2010.
lymphocytes into the peripheral blood compartment in human subjects. J. Appl. Ursolic acid overcomes Bcl-2-mediated resistance to apoptosis in prostate
Physiol. (1985) 103, 396–401. cancer cells involving activation of JNK-induced Bcl-2 phosphorylation and
Simpson, R.J., Guy, K., 2010. Coupling aging immunity with a sedentary lifestyle: has degradation. J. Cell. Biochem. 109, 764–773.
the damage already been done? – a mini-review. Gerontology 56, 449–458.

Downloaded for dr laqif abdurrahman (laqif.abdurrahman@gmail.com) at Universitas Indonesia from ClinicalKey.com by Elsevier on May 20, 2018.
For personal use only. No other uses without permission. Copyright ©2018. Elsevier Inc. All rights reserved.

You might also like