Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 Apr 23.
Published in final edited form as: Interdiscip Top Gerontol Geriatr. 2020 Apr 9;43:1–17. doi: 10.1159/000504480

How inflammation blunts innate immunity in aging

Emily L Goldberg 1, Albert C Shaw 2, Ruth R Montgomery 2,*
PMCID: PMC8063508  NIHMSID: NIHMS1692332  PMID: 32294641

Abstract

The collective loss of immune protection during aging leads to poor vaccine responses and an increased severity of infection for the elderly. Here we review our current understanding of effects of aging on the cellular and molecular dysregulation of innate immune cells as well as the relevant tissue milieu which influences their functions. The innate immune system is composed of multiple cell types which provide distinct and essential roles in tissue surveillance and antigen presentation as well as early responses to infection or injury. Functional defects that arise during aging lead to a reduced dynamic range of responsiveness, altered cytokine dynamics, and impaired tissue repair. Heightened inflammation influences both the dysregulation of innate immune responses as well as surrounding tissue microenvironments which have a critical role in development of a functional immune response. In particular, age-related physical and inflammatory changes in the skin, lung, lymph nodes, and adipose tissue reflect disrupted architecture and spatial organization contributing to diminished immune responsiveness. Underlying mechanisms include altered transcriptional programming and dysregulation of critical innate immune signaling cascades. Further, we identify signaling functions of bioactive lipid mediators which address chronic inflammation and may contribute to the resolution of inflammation to improve innate immunity during aging.

Keywords: monocyte, neutrophil, macrophage, dendritic cell, Toll-like receptor, bioactive lipid mediators, microenvironment, adipose tissue

INFLAMMATION AND AGING

Aging is the single greatest risk factor for developing chronic disease. Compared to younger adults, the elderly are at increased risk of infection and experience greater morbidity and mortality upon infection [1]. Similarly, the elderly also have poor vaccine responses, making them more susceptible to vaccine-preventable disease such as influenza or varicella zoster (causative agent in chickenpox and shingles) viruses, even when they have been vaccinated [2]. Central to this susceptibility is the functional deterioration of systemic immunity, or “immune senescence.” This term encompasses the collective loss of immune protection during aging and includes atrophy of the thymus leading to decreased naïve T cell output, an increased proportion of experienced lymphocytes contributing to adaptive immune memory but limiting responses to novel targets, an increased proportion of myeloid cells released from the bone marrow, and impaired functions of multiple existing immune cell types. Immune senescence is also characterized by functional dysregulation in immune cells at the cellular and molecular levels, ultimately leading to increased infection susceptibility and poor vaccination responses in the elderly as discussed in detail throughout this chapter.

Underlying many deficiencies in immunity is a chronic pro-inflammatory state. Many age-related diseases are driven by and promote increased inflammation, and elevated circulating inflammatory markers such as CRP, TNFα, and IL-6 correlate with age-related conditions including atherosclerosis [3] and frailty [4]. Importantly, even clinically healthy elderly individuals exhibit elevated inflammatory cytokines in serum [5]. While the precise source and cause of inflammation during aging is still unknown, this so-called “inflammaging” state contributes to the onset of multiple comorbidities in the elderly, including but not limited to Alzheimer’s disease, cardiovascular disease, muscle wasting, and immune senescence [6], making older adults a uniquely-challenging patient population. Here we will focus on the impact of inflammation on dysregulation of innate immune responses and relevant tissue microenvironments that influence immune cellular function.

CRITICAL FUNCTIONS OF INNATE IMMUNITY

The innate immune system is responsible for initial control of pathogens by directly eliminating infections, engaging nearby cells, and recruiting adaptive immune cells. In contrast to the adaptive immune system, which requires training and is exquisitely specific to particular pathogens, the innate immune system senses and responds to numerous infections directly. The innate immune system is composed of multiple cell types with distinct critical functions both in the circulation and following infiltration into tissues to provide early responses to infection or injury. Innate immune cells such as neutrophils and monocytes circulate throughout the body and are capable of rapidly infiltrating tissues; macrophages and dendritic cells (DCs) reside within tissue and have important roles in tissue surveillance and antigen presentation. Innate immune cells rely on a variety of pattern recognition receptors (PRRs) to sense tissue disturbance and also recognize pathogenic invasion. Membrane-associated Toll-like receptors (TLRs) on the cell surface and within endosomes sense diverse structural patterns associated with pathogens such as lipoproteins, lipopolysaccharide, flagellin, and nucleic acids. Numerous studies have reported diminished responsiveness to TLR ligands in innate immune cells from elderly donors as compared to younger donors and are discussed in more detail in the sections below. Nod-like receptors (NLRs) and RIG-I-like receptors (RLRs) are cytosolic sensors that respond to bacterial or viral molecular components to initiate inflammatory responses that limit pathogen spread. While aging affects each of these cell types and functional responses differently, some unifying mechanistic alterations offer insights to immune dysfunction in aging.

Neutrophils

Peripheral blood polymorphonuclear leukocytes (PMNs), the most abundant circulating white blood cells, account for 45–75% of circulating leukocytes. An estimated 1011 PMNs are released from the bone marrow daily to maintain a continuous supply of these crucial yet short-lived terminally differentiated cells. PMNs circulate throughout the body and therefore can potentially impact every tissue. They are rapid early responders to sites of infection or tissue injury and have high phagocytic and inflammatory capacity to limit pathogen spread. PMNs also contribute to chronic sterile inflammatory diseases such as gout in which they periodically accumulate and re-activate in afflicted joints causing debilitating pain in patients [7].

Despite their abundance and high inflammatory capacity, PMN are less well characterized in the field of aging immunology. PMN from older donors have lower TLR1 expression that correlates with reduced activation of TLR1-dependent IL-8, CD11b, and glucose uptake [8]. While PMN from healthy elderly donors have reduced phagocytic (FITC-labeled E. coli) capacity and increased superoxide in response to fMLP and PMA [9, 10], generalization of these findings has been complicated by differences noted from different stimuli and experimental conditions [11]. PMN from elderly individuals also show reduced actin polymerization [12, 13], suggesting impairments in chemotaxis. Indeed, in support of this possibility, PMN isolated from aged mouse bone marrow exhibited reduced chemotaxis [14]. In the aged, poor chemotaxis is proposed to prolong PMN presence in tissue, causing collateral tissue damage [14, 15].

PMN can undergo a novel form of cell death, NETosis, in which DNA content containing digestive and inflammatory enzymes are extruded from the cell. NETosis is a unique method to control pathogenic spread, but is also recently implicated in sterile inflammation [1618]. PMN from old mice have impaired ability to undergo NETosis in response to in vivo cecal ligation and puncture-induced model of sepsis and also in vitro after stimulation with TLR2 ligands, suggesting a cell-intrinsic defect in signaling to induce NET formation and/or extrusion [19, 20]. In parallel with these impairments, however, human PMNs from healthy older donors maintain their ability to activate the NLRP3 inflammasome when stimulated in vitro [21]. Despite certain PMN functions being retained during aging, the accumulated defects that have been identified outnumber them, and these multiple defects in critical early responding cells allow more rapid pathogenic spread early after infection, putting the elderly host at increased susceptibility to infection and morbidity.

Monocytes

Monocytes are a heterogeneous subset of circulating myeloid cells that can infiltrate tissues and differentiate into macrophages or dendritic cells. Their normal functions include phagocytosis, antigen presentation, and cytokine production. Multiple subsets of monocytes can be found in human blood at different stages of differentiation and maturity that are distinguishable by CD14 and CD16 expression [22]. Monocytes from older subjects have reduced production of cytokines after TLR1/2 stimulation that was associated with reduced surface TLR1 expression [23]; a generalized alteration in TLR-induced CD80 and CD86 expression correlates with reduced responses to influenza vaccination [24]. Monocytes from older subjects also have significantly diminished IFN-α/β responses to RIG-I stimulation [25]. Interestingly, these same monocytes retain the ability to produce pro-inflammatory cytokines upon stimulation, suggesting that aging may lead to cell-intrinsic dysregulation specifically in the IFN arm of this response [25, 26]. As there is no evidence of altered basal IFN expression with age, impaired IFN induction is representative of a model of age-related reduced dynamic range distinct from that of TLR-mediated proinflammatory cytokine induction. However, a significantly higher percentage of unstimulated monocytes from older donors exhibited nuclear NF-κB (p65) translocation, i.e. a higher activation status at baseline, and these cells secreted significantly more TLR5-induced IL-8 compared to monocytes from younger individuals [27], representing a possible avenue for vaccine adjuvant design. The age-related defects in monocytes have not yet been reported in other innate immune cells and highlight that each cell subset accumulates its own functional defects ultimately culminating in impaired innate immune protection during aging.

Macrophages

Macrophages are versatile innate immune cells that are important for initiating pro-inflammatory immune responses in addition to roles in phagocytosis, resolution, and tissue repair after injury. In the steady-state, anti-inflammatory macrophages help maintain homeostatic conditions within the tissue and become activated in response to infection or injury. Activated macrophages can secrete a variety of cytokines (e.g. TNFα, IL-1β, NO) that prime the inflammatory immune response and chemokines (MIP-2, KC) that recruit additional immune cells. After pathogen clearance, anti-inflammatory macrophages help clear apoptotic and necrotic debris and secrete cytokines and growth factors that promote tissue repair and regeneration. Mouse studies have shown the large variety of functions carried out by macrophages reflects the existence of multiple subsets and lineages of macrophages, each occupying distinct niches within tissues. For example, resident macrophages have recently been reported to interact directly with the nervous system in the intestine [28] and adipose tissue [29, 30] to regulate tissue function. Taken together, these data indicate macrophages have a central role in maintaining tissue homeostasis and also orchestrating immune responses.

Despite this importance, relatively little is known about changes in tissue macrophage function during aging, largely due to their poor accessibility within tissues. Monocyte-derived macrophages from human elderly human blood donors show impaired DC-SIGN-induced reduction in the expression of TLR3 following infection with West Nile virus (WNV) in vitro [31]. This impairment via the signal transducer and activator of transcription 1 (STAT1)-mediated pathway may be relevant for elevated cytokine production contributing to permeability of the blood-brain barrier and increased severity of WNV infection in older individuals [31, 32]. Murine studies also support an age-related increase in Influenza A virus (IAV) susceptibility that may contribute to impaired tissue repair and altered cytokine dynamics [3335]. In aged mice, elicited peritoneal macrophages showed reduced phagocytosis of FITC-labeled E. coli [36]. Importantly, this same study found in vitro differentiated bone marrow-derived macrophages from old mice did not have this defect, suggesting mechanisms beyond cell-intrinsic changes during aging. Notably, peritoneal macrophages transferred from adult into old mice lost their phagocytic capacity. This result emphasizes that the tissue environment is a critical determinant of macrophage function.

Dendritic Cells

DCs are key sources of inflammatory cytokines and co-stimulatory molecules that instruct the development of the adaptive anti-microbial immune response. Multiple classes of DCs exist, with the most common being myeloid DC (mDC) that activate naïve T cells, and plasmacytoid DC (pDC) which are major sources of IFNα following viral infection. DCs reside within lymph nodes or other tissues and survey the local microenvironment and then migrate to nearby lymph nodes that are being patrolled by adaptive immune cells. Recent studies using a unique resource of tissue acquisition from human organ donors has revealed that DC subset composition varies by tissue and age in humans, and these changes may impact site-specific immunity [37]. Both mDC and pDC from older donors show lower expression of TLRs globally and substantial decreases in cytokine production following TLR stimulation [38, 39]. Recent studies of DCs detected lower levels of RIG-I from older human subjects [40]. Similarly, pDCs and monocyte-derived DCs from healthy older subjects also secrete less IFN in response to IAV [4143] and to WNV [38, 40]. DC production of type I IFN was significantly lower in older donors compared to younger donors, with diminished induction of late phase signaling responses, e.g. STAT1, IRF7, and IRF1 suggesting defective regulation of type I IFN induction [40]. IFN production by pDCs is decreased in older HSV-2 infected mice owing to impaired IRF7 upregulation upon viral infection, potentially further compromising antiviral immunity [44]. Multiple functional defects in these critical cells that bridge the innate and adaptive immune responses greatly contribute to impaired immune activation and responsiveness to infection in the elderly.

INFLAMMATION AND INNATE IMMUNITY

While most vaccine responses are assessed by generation of antigen-specific memory T cells and protective antibody titers, proper activation of the innate immune system is essential for orchestrating this process. Innate immune cells create local cytokine and chemokine gradients to recruit cells to sites of damage or infection. They also carry antigen back to lymph nodes to initiate adaptive immune cell activation and expansion. During aging many of these critical functions performed by innate immune cells become compromised, leading to poor overall immune protection and vaccine responsiveness. The chronically elevated basal levels of pro-inflammatory cytokines in the elderly lead to impaired responsiveness of innate immune cells by raising the threshold of activation, thereby compressing the dynamic range of responsiveness. Evidence from multiple studies in old mice illustrate the role of inflammatory cytokines in impaired innate immune function during aging. Splenic macrophages isolated from old IL-6-deficient mice have restored secretion of TNFα, IL-1β, and IL-12 after ex vivo stimulation with LPS [45]. Similarly, inflammatory monocytes were found to correlate with elevated serum TNFα and interfere with bacterial clearance from the lungs of old mice, but aged TNFα-deficient mice were able to effectively clear the infection [46]. Taken together, these data support a hypothesis that continuous bathing of innate immune cells in the inflammatory milieu of the aged individual reduces the abilities of these cells to sense and respond to signals such as tissue damage, infection, or vaccination.

Influence of tissue milieu on innate immune cell function

Much of what is known about innate immune cell functional changes during aging has been elucidated in vitro in cells isolated from peripheral blood and do not reflect tissue milieu. Mounting a functional immune response depends not only on intrinsic responses by innate immune cells, but also their ability to communicate with the neighboring cells around them. Animal models and studies of ex vivo human tissues have provided some insights into how the tissue microenvironment significantly shapes the function and identity of its resident immune cells [37, 4749]. Rodent studies utilizing heterochronic parabiosis, the surgical joining of a young and old animal in which a shared circulatory system develops, have revealed environmental defects in the aged animal that can be improved by exposure to circulating factors from the young animal [50]. Similarly, adoptive cell transfers of adult or old cells into reciprocally-aged hosts reveal that the aged environment impairs functional responses of innate immune cells [36, 51, 52]. Although these experimental techniques are limited to inbred animal models, innovative studies in lung transplant patients have recently been used to study tissue-resident T cells in humans [53]. In the following sections we describe selected tissues that exhibit strong age-related alterations and discuss how these might impact innate immunity and vaccine efficacy.

Skin

The skin is the largest organ in the human body and represents one of the first physical barriers to protect against pathogens. Skin also contains a unique composition of innate immune cells, including macrophages, innate-like γδ T cells, as well as classical memory αβ T cells, and a specialized subset of DCs called Langerhans cells. While obvious physical appearance reflects important cellular and tissue organizational changes in aged skin, we lack a detailed understanding of how the immune system in skin changes with age. To address this gap, sophisticated methods have been developed by Dr. Arne Akbar and colleagues in which a suction blister is used to collect leukocytes and fluid from the skin after cutaneous antigen challenge. They found impaired memory T cell migration to the skin after challenge with fungal (Candida albicans), viral (Varicella zoster virus, VZV), and mycobacterial (tuberculin PPD) antigen challenges. They further showed with the C. albicans model that impaired memory CD4 T cell homing to aged skin is due to reduced TNFα secretion from macrophages [54]. The reduced TNFα led to impaired endothelial activation of selectin molecules including E-selectin, VCAM-1, and ICAM-1. Ultimately the lower adhesion molecule expression on the endothelium led to reduced memory CD4 T cell migration to the skin. Importantly, the defects in both the endothelium and macrophages could be restored in vitro, suggesting they reflect the influence of the tissue environment within the skin and may be reversible.

Skin biopsies from elderly individuals exhibit elevated p38 MAPK transcriptional signatures that correlated with impaired VZV skin response. Treating these subjects with an oral MAPK inhibitor prior to antigen challenge improves systemic inflammation and their VZV-specific recall response in the skin [55]. Similar depressed immune responses have been reported in a murine skin infection model of Staphylococcus aureus in which old mice have delayed bacterial clearance, delayed wound closure, and reduced neutrophil chemotaxis to the wound site [56]. This highlights the detrimental effects of elevated basal inflammation that compresses the dynamic range of innate immune activation in the elderly. Improving immunity in aging skin has potentially enormous clinical implications particularly as skin is a common vaccination site. Perhaps improving immune cell recruitment to and from the skin and draining lymph nodes represents a realistic approach towards achieving the goal of improved vaccine efficacy in the elderly.

Lung

Like the skin, the lung is another barrier tissue constantly exposed to environmental and microbial pathogens. Individuals aged 75–84 have nearly 20-times higher morbidity and mortality compared to younger adults after acute lung injury [57]. People over age 65 account for 70% of influenza- and pneumonia-related hospitalizations [58] and studies in mice show that aging increases susceptibility to secondary bacterial challenges after influenza infection [59]. Together these studies highlight multiple defects in the aging lung, including local immune responsiveness and tissue maintenance and repair, which increases host vulnerability to lung damage and disease. Age-related changes in the lung milieu are significant given that even in young animals the lung environment strongly regulates the activation and function of resident immune cells [60]. In a prospective study of emergency room patients with burn inhalation injuries, older patients were at increased risk of death after burn injury, and had higher concentrations of inflammatory cytokines in their serum and bronchoalveolar lavage fluid [61]. High vascularization within the lung makes it a unique site that can be rapidly infiltrated by circulating leukocytes. Increased neutrophil infiltration and activation is associated with excess immune pathology during infection in mouse and human studies [25, 6266]. As described above, dysregulation of neutrophil responses and impaired chemotaxis could impair pathogen clearance and prolong detrimental inflammatory responses in the lungs of elderly individuals.

Lymph node

Lymph nodes are the central hubs in which innate and adaptive immunity coordinate productive immune responses. They are highly-organized structures strategically placed at intersections between draining lymphatics and the circulating blood vasculature. Upon encountering and processing pathogens in the periphery, innate immune cells such as DCs migrate to nearby lymph nodes. Migration of cells towards and within lymph nodes has been expertly reviewed previously [67]; resident innate immune cells, including macrophages, are poised to respond immediately in the case of pathogen entry, or relay immune information further into the lymph node. T cell and B cell zones in the inner cortex are spatially organized and surrounded by macrophages. Lymphocytes enter through high endothelial venules (HEVs), guided by fibroblastic reticular cells. Here they scan migratory DCs for cognate antigen. T and B cells that are continuously circulating and patrolling the body migrate through these lymph nodes and upon encountering a DC-presented cognate antigen, clonally expand and mount a massive and highly-specific adaptive immune response with the purpose of eradicating the microbial pathogen and generating long-lasting immunity. In order for this entire process to be successful, many coordinated events must be executed successfully.

Proper function of lymph nodes requires intense cross-talk between hematopoietic cells and the stroma. Accumulation of lipid droplets and increased fibrosis are common features of the disrupted lymph node architecture and disrupted spatial organization associated with aging [68, 69]. These structural changes not only impair physical communication between lymphocytes and the lymph node stroma, but also probably perturb the environment and disrupt normal homeostasis within the lymph node. Multiple studies have reported reduced lymphocyte cellularity of aged lymph nodes both during the basal state and during infection [52, 68, 70]. Disrupted organization of T cell and B cell zones in the aging lymph node further compounds the diminished immune response and may explain, in part, lower-magnitude T cell responses and lower antibody responses after infection or vaccination in the elderly.

Adipose tissue

Although classically considered an energy-storage organ, it is becoming increasingly clear that adipose tissue is also an immunologically responsive organ that contributes to systemic inflammation. Adipose tissue remodeling and redistribution into abdominal fat are common features of age-related adipose tissue dysfunction. Importantly, these physiological changes can occur independently of obesity. In addition to increased visceral adiposity, ectopic lipid accumulation in tissues, including bone marrow, the thymus, liver, and muscle increases during healthy aging and can upset normal tissue homeostasis [71].

As adipose tissue inappropriately accumulates in tissues and lymphoid organs it disrupts tissue architecture, function, and perhaps essential cell-cell communication. Adipose tissue secretes cytokine-like molecules--known as adipokines--such as leptin and adiponectin that modulate immune cell function. Besides these secreted factors, the adipose tissue itself contains a unique immune phenotype [72]. Under steady-state conditions the adipose contains primarily anti-inflammatory macrophages, γδ T cells, and other innate immune cells with tissue maintenance and reparative properties. However, during aging, the composition of the adipose-resident immune populations become skewed and heavily enriched for pro-inflammatory macrophages, B cells, and memory T cells [7375].

Mouse studies indicate that adipose-resident macrophages exhibit particularly intriguing changes during aging. Unlike obesity, in which pro-inflammatory macrophages increase numerically, aging is accompanied by an overall reduction in the proportion of macrophages in visceral adipose tissue, although the population still manifests a generally pro-inflammatory phenotype [73]. The macrophages that remain occupy several distinct niches, including crown-like structures surrounding dead or dying adipocytes, scattered in the parenchyma, and lining sympathetic nerves within the adipose tissue [29]. The “aged” macrophages gain a unique transcriptional profile that includes upregulation of enzymes monoamine oxidase A (Maoa) and catechol-O-methyltransferase (Comt) that degrade catecholamines, leading to impaired lipolysis during fasting. This phenotype may also be driven by chronic low-grade inflammation by macrophages, as NLRP3 inflammasome-deficient old mice are protected from lipolysis resistance. Notably, NLRP3-deficient mice are also protected from numerous age-related inflammatory diseases, including aspects of immunosenescence [76, 77],sarcopenia [78], and experimental lung fibrosis [79] highlighting its role as a potential driver of age-related inflammation. While human studies show transcriptional activation of inflammasome gene signatures during aging [78], more studies are needed to formally understand how these pro-inflammatory immune complexes promote immune senescence during aging.

MECHANISMS OF DYSFUNCTION IN INNATE IMMUNE CELLS DURING AGING

For the numerous innate immune defects that develop with age, the impact of tissue milieu on innate immune cells may have particular relevance, but data from humans is limited, and additional novel approaches are desperately needed. Multiple mechanisms that cause cell-intrinsic defects have been identified in experiments using innate immune cells isolated from peripheral blood of elderly and adult donors. Notably, these mechanisms may occur simultaneously within particular cells, or across multiple populations, ultimately disarming the proper function of the aged innate immune system.

Altered transcription after stimulation

Transcriptional programming coordinates the pro-inflammatory programs needed for innate immune control of pathogens. PBMC from healthy aged participants had elevated baseline type 1 interferon gene signature compared to PBMC from younger adults, and levels could not be further induced by influenza vaccination [80]. Younger adult PBMC also showed increased oxidative phosphorylation and mitochondrial biogenesis programs that were absent in older adult PBMC, particularly in older adults that did not respond to vaccination. Circulating isolated monocyte subsets from elderly individuals showed altered transcriptional profiles in response to ex vivo TLR ligand stimulation including exaggerated superoxide and oxidative stress program signatures [81]. A large multi-cohort analysis showed that baseline gene expression predicts influenza vaccination response in young adults. Fifteen genes (up-regulated: RAB24, GRB2, DPP3, ACTB, MVP, DPP7, ARPC4, PLEKHB2, ARRB1; down-regulated: PTPN22, PURA, SP4, CASP6, NUDCD2) were identified in young adults classified as “high responders” to influenza vaccination. These nine upregulated genes were not induced in older individuals and no genes in older adults were differentially expressed in low and high responders, suggesting distinct responses lead to lower vaccine response in older individuals [82]. Hematopoietic stem cells (HSCs) from older individuals exhibit hypermethylation of the IRF8 locus [83], and thus epigenetic regulatory mechanisms may link the well-known myeloid skewing of HSC with impaired IFN induction with age [26]. These age-related impairments may contribute to impaired vaccine response and the increased susceptibility of older persons to IAV infection.

Impaired Mer signaling

TAM receptors (Tyro3, Axl, and Mer) are a family of receptor tyrosine kinases that play critical roles in tissue homeostasis by recognizing and inducing phagocytosis of apoptotic cells. TAMs are also negative regulators of TLR-mediated immune responses that broadly inhibit both TLR and TLR-induced cytokine receptor cascades to limit inflammation [84]. The importance of TAMs in immune activation is illustrated by the observation that reduced levels of TAMs in humans and in a TAM-deficient mouse model are associated with susceptibility to autoimmune disease and higher or chronic inflammation [8587]. Monocytes from older donors showed elevated expression of TAM receptors but impaired activation of the Mer pathway following binding of the ligand Protein S, leading to impaired signaling through AKT [88]. The elevated expression of TAM receptors in monocytes from older adults has important implications for dysregulation of immune responses in aging--in particular as the Mer pathway is critical for clearance of apoptotic cells that contribute to inflammation in aging [89, 90]. Defective TAM signaling in alveolar macrophages in old mice may also explain impaired phagocytic clearance of apoptotic PMN and increased mortality during influenza infection [91]. Further, TAMs play a key role in mediating autophagy [84, 85], and the reduced efficiency of autophagy in aging has been shown to contribute to accumulation of damaged proteins in cells [92]. The importance of TAMs in aging may be especially significant in tissue, where levels of Mer are highest, and may present avenues for modulation of chronic tissue inflammation noted in aging.

Reduced RIG-I signaling

Innate immune cells rely on intracellular sensors known as PRRs to recognize pathogen-associated molecular patterns (PAMPs). One such PRR induced during viral infection by recognition of 5’-triphosphate (5’-ppp) is the retinoic acid-inducible gene-I (RIG-I) which induces Type 1 IFN to control viral infections. Recent studies of DCs detected lower levels of RIG-I from older human subjects [40] and monocytes from older subjects have significantly diminished IFN-α/β responses to RIG-I stimulation [25]. Total human PBMCs exhibit impaired IFN responses to 5’-ppp RNA transfection relative to younger controls 6 hours after stimulation, although in this complex cell population responses were comparable after 24 hours [93]. Studies of potential mechanisms mediating these age-associated findings revealed that monocytes from older adults exhibit decreased expression of the adaptor protein TRAF3 as a consequence of its increased proteasomal degradation with age, thereby impairing the RIG-I primary signaling pathway for type I IFNs. Monocytes from older adults also fail to effectively upregulate the interferon regulatory transcription factor IRF8, compromising their ability to participate in IFN induction through secondary RIG-I signaling [26]. Such RIG-I signaling defects in multiple cell types could likely contribute to increased risk for infection and morbidity and mortality from viral infections in the context of aging. As the innate immune system helps instruct appropriate responses of the adaptive immune system, these innate signaling defects may also contribute to impaired adaptive immunity and vaccine efficacy during aging.

Energy Balance and Inflammation Regulation

Recent progress targeting metabolic programming suggests potential treatment strategies to enhance immune responses in older adults. While several non-pharmaceutical dietary interventions have demonstrated improved longevity in animal models, including calorie restriction, fasting-mimicking diet, and ketogenic diet, few have investigated their impact on immune function during aging [9497], and this area presents opportunities for future investigation. Many lifespan-extending interventions induce a negative energy balance in which energy expenditure exceeds energy consumption [98101]. In this adaptive starvation response, the body switches from glucose to lipid breakdown to support energetic requirements via production of short-chain fatty acids such as ketone bodies. Accordingly, ketogenic diets comprised almost entirely of fat and protein actually promote fat break down and weight loss because of this adaptation to low-glucose availability. In addition to serving as an alternative fuel source, the most abundant ketone body, β-hydroxybutyrate, potently inhibits NLRP3 inflammasome activation in macrophages, monocytes, and neutrophils from adult and older humans and mice [21, 102] and extends lifespan in C. elegans [103]. Notably, two studies recently reported increased lifespan and improved cognition in old mice fed a ketogenic diet [95, 96]. This highlights the importance of whole-body metabolism in regulating age-related inflammation and presents a critical window for future exploration.

Immune Cell Energy Metabolism

Metabolic programming in immune cells is a critical determinant of their downstream functions and mounting a robust immune response is energetically expensive. Increased glycolysis is generally associated with elevated pro-inflammatory activity whereas fatty acid oxidation and oxidative phosphorylation is associated with quiescence and anti-inflammatory functions [104]. At the population level, a balance between these programs allows for pro-inflammatory immunemediated pathogen clearance, while preserving tissue repair functions and long-lived immunity in the case of adaptive immune cells. In addition to the importance of overall metabolic programming, specific metabolites such as succinate, itaconate, and β-hydroxybutyrate regulate immune cell function [105, 106]. Therefore, targeting immune cell metabolism represents at attractive strategy for modulating immune responsiveness and vaccine efficacy. It should be noted that the simplified model presented here is derived from mostly in vitro studies and whether this programming balance occurs in vivo, or is maintained during aging is not known. Future studies are needed to test whether basal metabolic programming and/or activation-induced metabolic upregulation in innate immune cells is retained during aging and what cues stimulate this reprogramming.

Bioactive Lipids

In addition to metabolic programming, dietary metabolites can have important cell signaling functions. While we have focused on innate immune activation defects that contribute to increased infection susceptibility in the elderly, an equally important aspect of the innate immune system is the resolution of inflammation. Key to resolution is a family of pro-resolution bioactive lipid mediators synthesized from omega-3 fatty acids, namely docosahexanoic acid (DHA) and eicosapentaenoic acid (EPA)---the resolvins, lipoxins, and maresins [107]. Aged mice exhibit delayed bacterial clearance and have reduced levels of these pro-resolving lipid metabolites [108]. Innate immune cells are capable of metabolizing DHA and EPA to generate these metabolites, and this improves efferocytosis in human monocytes [108]. Finally, MerTK-dependent ERK activation induces pro-resolving lipoxin LXA4 production in human monocyte-derived macrophages [109]. Perhaps age-related changes in tissue milieu can be restored by improving inflammation resolution in the elderly. Future investigations in this domain are warranted to explore their potential for dampening chronic inflammation during aging to improve innate immunity.

PERSPECTIVES AND FUTURE DIRECTIONS

The impact of chronic inflammation on innate immunity, both within immune cells and their tissue microenvironment, highlight a critical area for future investigation and potential novel therapeutic interventions. As we have described, multiple interconnected deficits arise during aging, both within innate immune cell subsets and the tissue milieu, to impair in vivo innate immunity in the elderly. With these complex layers of regulation, it remains challenging to fully define mechanisms of innate immune deficiencies. Collectively these defects lead to poor priming of the adaptive immune system, culminating in poor immune protection and vaccination in the elderly. Given that in vitro stimulation of isolated circulating immune cells probably does not fully reflect their responsiveness within the tissue milieu, future studies will require appropriate in situ tissue studies and animal models. Along with development of new approaches to define age-related defects in innate immunity, a critical priority is novel therapeutic interventions targeting systemic inflammation to address immune dysfunction and enhance healthy life-span.

ACKNOWLEDGEMENTS

The authors wish to thank their research groups and members of the Yale Center for Research on Aging for helpful discussions and many colleagues whose work could not be cited. This work was supported in part by grants from the NIH (AI 089992 and AG055362 to RRM and ACS, and K99-AG058801 to ELG).

REFERENCES

  • 1.Liang SY and Mackowiak PA, Infections in the elderly. Clin Geriatr Med 2007. 23: 441–456, viii. [DOI] [PubMed] [Google Scholar]
  • 2.Haq K and McElhaney JE, Immunosenescence: Influenza vaccination and the elderly. Curr Opin Immunol 2014. 29: 38–42. [DOI] [PubMed] [Google Scholar]
  • 3.Shen-Orr SS, Furman D, Kidd BA, Hadad F, Lovelace P, Huang YW, Rosenberg-Hasson Y, Mackey S, Grisar FA, Pickman Y, Maecker HT, Chien YH, Dekker CL, Wu JC, Butte AJ and Davis MM, Defective Signaling in the JAK-STAT Pathway Tracks with Chronic Inflammation and Cardiovascular Risk in Aging Humans. Cell Syst 2016. 3: 374–384 e374. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Marcos-Perez D, Sanchez-Flores M, Maseda A, Lorenzo-Lopez L, Millan-Calenti JC, Gostner JM, Fuchs D, Pasaro E, Laffon B and Valdiglesias V, Frailty in Older Adults Is Associated With Plasma Concentrations of Inflammatory Mediators but Not With Lymphocyte Subpopulations. Front Immunol 2018. 9: 1056. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Tanaka T, Biancotto A, Moaddel R, Moore AZ, Gonzalez-Freire M, Aon MA, Candia J, Zhang P, Cheung F, Fantoni G, consortium CHI, Semba RD and Ferrucci L, Plasma proteomic signature of age in healthy humans. Aging Cell 2018: e12799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Franceschi C, Bonafe M, Valensin S, Olivieri F, De Luca M, Ottaviani E and De Benedictis G, Inflamm-aging. An evolutionary perspective on immunosenescence. Ann N Y Acad Sci 2000. 908: 244–254. [DOI] [PubMed] [Google Scholar]
  • 7.Dalbeth N, Merriman TR and Stamp LK, Gout. Lancet 2016. [DOI] [PubMed] [Google Scholar]
  • 8.Qian F, Guo X, Wang X, Yuan X, Chen S, Malawista SE, Bockenstedt LK, Allore HG and Montgomery RR, Reduced bioenergetics and toll-like receptor 1 function in human polymorphonuclear leukocytes in aging. Aging (Albany NY) 2014. 6: 131–139. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Butcher SK, Chahal H, Nayak L, Sinclair A, Henriquez NV, Sapey E, O’Mahony D and Lord JM, Senescence in innate immune responses: reduced neutrophil phagocytic capacity and CD16 expression in elderly humans. J Leukoc Biol 2001. 70: 881–886. [PubMed] [Google Scholar]
  • 10.Braga PC, Sala MT, Dal Sasso M, Pecile A, Annoni G and Vergani C, Age-associated differences in neutrophil oxidative burst (chemiluminescence). Exp Gerontol 1998. 33: 477–484. [DOI] [PubMed] [Google Scholar]
  • 11.Tortorella C, Piazzolla G, Spaccavento F, Vella F, Pace L and Antonaci S, Regulatory role of extracellular matrix proteins in neutrophil respiratory burst during aging. Mech Ageing Dev 2000. 119: 69–82. [DOI] [PubMed] [Google Scholar]
  • 12.Rao KM, Age-related decline in ligand-induced actin polymerization in human leukocytes and platelets. J Gerontol 1986. 41: 561–566. [DOI] [PubMed] [Google Scholar]
  • 13.Rao KM, Currie MS, Padmanabhan J and Cohen HJ, Age-related alterations in actin cytoskeleton and receptor expression in human leukocytes. J Gerontol 1992. 47: B37–44. [DOI] [PubMed] [Google Scholar]
  • 14.Sapey E, Greenwood H, Walton G, Mann E, Love A, Aaronson N, Insall RH, Stockley RA and Lord JM, Phosphoinositide 3-kinase inhibition restores neutrophil accuracy in the elderly: toward targeted treatments for immunosenescence. Blood 2014. 123: 239–248. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Montgomery RR and Shaw AC, Paradoxical changes in innate immunity in aging: recent progress and new directions. J Leukoc Biol 2015. 98: 937–943. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Jorch SK and Kubes P, An emerging role for neutrophil extracellular traps in noninfectious disease. Nat Med 2017. 23: 279–287. [DOI] [PubMed] [Google Scholar]
  • 17.Schauer C, Janko C, Munoz LE, Zhao Y, Kienhofer D, Frey B, Lell M, Manger B, Rech J, Naschberger E, Holmdahl R, Krenn V, Harrer T, Jeremic I, Bilyy R, Schett G, Hoffmann M and Herrmann M, Aggregated neutrophil extracellular traps limit inflammation by degrading cytokines and chemokines. Nat Med 2014. 20: 511–517. [DOI] [PubMed] [Google Scholar]
  • 18.Fadini GP, Menegazzo L, Rigato M, Scattolini V, Poncina N, Bruttocao A, Ciciliot S, Mammano F, Ciubotaru CD, Brocco E, Marescotti MC, Cappellari R, Arrigoni G, Millioni R, Vigili de Kreutzenberg S, Albiero M and Avogaro A, NETosis Delays Diabetic Wound Healing in Mice and Humans. Diabetes 2016. 65: 1061–1071. [DOI] [PubMed] [Google Scholar]
  • 19.Xu F, Zhang C, Zou Z, Fan EKY, Chen L, Li Y, Billiar TR, Wilson MA, Shi X and Fan J, Aging-related Atg5 defect impairs neutrophil extracellular traps formation. Immunology 2017. 151: 417–432. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Hazeldine J, Harris P, Chapple IL, Grant M, Greenwood H, Livesey A, Sapey E and Lord JM, Impaired neutrophil extracellular trap formation: a novel defect in the innate immune system of aged individuals. Aging Cell 2014. 13: 690–698. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Goldberg EL, Asher JL, Molony RD, Shaw AC, Zeiss CJ, Wang C, Morozova-Roche LA, Herzog RI, Iwasaki A and Dixit VD, beta-Hydroxybutyrate Deactivates Neutrophil NLRP3 Inflammasome to Relieve Gout Flares. Cell Rep 2017. 18: 2077–2087. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Ziegler-Heitbrock L, Blood Monocytes and Their Subsets: Established Features and Open Questions. Front Immunol 2015. 6: 423. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.van Duin D, Mohanty S, Thomas V, Ginter S, Montgomery RR, Fikrig E, Allore HG, Medzhitov R and Shaw AC, Age-associated defect in human TLR-1/2 function. J Immunol 2007. 178: 970–975. [DOI] [PubMed] [Google Scholar]
  • 24.van Duin D, Allore HG, Mohanty S, Ginter S, Newman FK, Belshe RB, Medzhitov R and Shaw AC, Prevaccine determination of the expression of costimulatory B7 molecules in activated monocytes predicts influenza vaccine responses in young and older adults. J Infect Dis 2007. 195: 1590–1597. [DOI] [PubMed] [Google Scholar]
  • 25.Pillai PS, Molony RD, Martinod K, Dong H, Pang IK, Tal MC, Solis AG, Bielecki P, Mohanty S, Trentalange M, Homer RJ, Flavell RA, Wagner DD, Montgomery RR, Shaw AC, Staeheli P and Iwasaki A, Mx1 reveals innate pathways to antiviral resistance and lethal influenza disease. Science 2016. 352: 463–466. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Molony RD, Nguyen JT, Kong Y, Montgomery RR, Shaw AC and Iwasaki A, Aging impairs both primary and secondary RIG-I signaling for interferon induction in human monocytes. Sci Signal 2017. 10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Qian F, Wang X, Zhang L, Chen S, Piecychna M, Allore H, Bockenstedt L, Malawista S, Bucala R, Shaw AC, Fikrig E and Montgomery RR, Age-associated elevation in TLR5 leads to increased inflammatory responses in the elderly. Aging Cell 2012. 11: 104–110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Gabanyi I, Muller PA, Feighery L, Oliveira TY, Costa-Pinto FA and Mucida D, Neuro-immune Interactions Drive Tissue Programming in Intestinal Macrophages. Cell 2016. 164: 378–391. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Camell CD, Sander J, Spadaro O, Lee A, Nguyen KY, Wing A, Goldberg EL, Youm YH, Brown CW, Elsworth J, Rodeheffer MS, Schultze JL and Deep Dixit V, Inflammasome-driven catecholamine catabolism in macrophages blunts lipolysis during ageing. Nature 2017. 550: 119–123. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Pirzgalska RM, Seixas E, Seidman JS, Link VM, Sanchez NM, Mahu I, Mendes R, Gres V, Kubasova N, Morris I, Arus BA, Larabee CM, Vasques M, Tortosa F, Sousa AL, Anandan S, Tranfield E, Hahn MK, Iannacone M, Spann NJ, Glass CK and Domingos AI, Sympathetic neuron-associated macrophages contribute to obesity by importing and metabolizing norepinephrine. Nat Med 2017. 23: 1309–1318. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Kong KF, Delroux K, Wang X, Qian F, Arjona A, Malawista SE, Fikrig E and Montgomery RR, Dysregulation of TLR3 impairs the innate immune response to West Nile virus in the elderly. J Virol 2008. 82: 7613–7623. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Hoffman KW, Sachs D, Bardina SV, Michlmayr D, Rodriguez CA, Sum J, Foster GA, Krysztof D, Stramer SL and Lim JK, Differences in Early Cytokine Production Are Associated With Development of a Greater Number of Symptoms Following West Nile Virus Infection. J Infect Dis 2016. 214: 634–643. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Yin L, Zheng D, Limmon GV, Leung NH, Xu S, Rajapakse JC, Yu H, Chow VT and Chen J, Aging exacerbates damage and delays repair of alveolar epithelia following influenza viral pneumonia. Respir Res 2014. 15: 116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Toapanta FR and Ross TM, Impaired immune responses in the lungs of aged mice following influenza infection. Respir Res 2009. 10: 112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Zhao J, Zhao J, Legge K and Perlman S, Age-related increases in PGD(2) expression impair respiratory DC migration, resulting in diminished T cell responses upon respiratory virus infection in mice. J Clin Invest 2011. 121: 4921–4930. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Linehan E, Dombrowski Y, Snoddy R, Fallon PG, Kissenpfennig A and Fitzgerald DC, Aging impairs peritoneal but not bone marrow-derived macrophage phagocytosis. Aging Cell 2014. 13: 699–708. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Granot T, Senda T, Carpenter DJ, Matsuoka N, Weiner J, Gordon CL, Miron M, Kumar BV, Griesemer A, Ho SH, Lerner H, Thome JJC, Connors T, Reizis B and Farber DL, Dendritic Cells Display Subset and Tissue-Specific Maturation Dynamics over Human Life. Immunity 2017. 46: 504–515. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Jing Y, Shaheen E, Drake RR, Chen N, Gravenstein S and Deng Y, Aging is associated with a numerical and functional decline in plasmacytoid dendritic cells, whereas myeloid dendritic cells are relatively unaltered in human peripheral blood. Hum Immunol 2009. 70: 777–784. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Panda A, Qian F, Mohanty S, van Duin D, Newman FK, Zhang L, Chen S, Towle V, Belshe RB, Fikrig E, Allore HG, Montgomery RR and Shaw AC, Age-associated decrease in TLR function in primary human dendritic cells predicts influenza vaccine response. J Immunol 2010. 184: 2518–2527. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Qian F, Wang X, Zhang L, Lin A, Zhao H, Fikrig E and Montgomery RR, Impaired interferon signaling in dendritic cells from older donors infected in vitro with West Nile virus. J Infect Dis 2011. 203: 1415–1424. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Sridharan A, Esposo M, Kaushal K, Tay J, Osann K, Agrawal S, Gupta S and Agrawal A, Age-associated impaired plasmacytoid dendritic cell functions lead to decreased CD4 and CD8 T cell immunity. Age (Dordr) 2011. 33: 363–376. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Canaday DH, Amponsah NA, Jones L, Tisch DJ, Hornick TR and Ramachandra L, Influenza-induced production of interferon-alpha is defective in geriatric individuals. J Clin Immunol 2010. 30: 373–383. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Prakash S, Agrawal S, Cao JN, Gupta S and Agrawal A, Impaired secretion of interferons by dendritic cells from aged subjects to influenza : role of histone modifications. Age (Dordr) 2013. 35: 1785–1797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Stout-Delgado HW, Yang X, Walker WE, Tesar BM and Goldstein DR, Aging impairs IFN regulatory factor 7 up-regulation in plasmacytoid dendritic cells during TLR9 activation. J Immunol 2008. 181: 6747–6756. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Gomez CR, Karavitis J, Palmer JL, Faunce DE, Ramirez L, Nomellini V and Kovacs EJ, Interleukin-6 contributes to age-related alteration of cytokine production by macrophages. Mediators Inflamm 2010. 2010: 475139. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Puchta A, Naidoo A, Verschoor CP, Loukov D, Thevaranjan N, Mandur TS, Nguyen PS, Jordana M, Loeb M, Xing Z, Kobzik L, Larche MJ and Bowdish DM, TNF Drives Monocyte Dysfunction with Age and Results in Impaired Anti-pneumococcal Immunity. PLoS Pathog 2016. 12: e1005368. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Thome JJ, Yudanin N, Ohmura Y, Kubota M, Grinshpun B, Sathaliyawala T, Kato T, Lerner H, Shen Y and Farber DL, Spatial map of human T cell compartmentalization and maintenance over decades of life. Cell 2014. 159: 814–828. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Lavin Y, Winter D, Blecher-Gonen R, David E, Keren-Shaul H, Merad M, Jung S and Amit I, Tissue-resident macrophage enhancer landscapes are shaped by the local microenvironment. Cell 2014. 159: 1312–1326. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Xue J, Schmidt SV, Sander J, Draffehn A, Krebs W, Quester I, De Nardo D, Gohel TD, Emde M, Schmidleithner L, Ganesan H, Nino-Castro A, Mallmann MR, Labzin L, Theis H, Kraut M, Beyer M, Latz E, Freeman TC, Ulas T and Schultze JL, Transcriptome-based network analysis reveals a spectrum model of human macrophage activation. Immunity 2014. 40: 274–288. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Conboy MJ, Conboy IM and Rando TA, Heterochronic parabiosis: historical perspective and methodological considerations for studies of aging and longevity. Aging Cell 2013. 12: 525–530. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Li G, Smithey MJ, Rudd BD and Nikolich-Zugich J, Age-associated alterations in CD8alpha+ dendritic cells impair CD8 T-cell expansion in response to an intracellular bacterium. Aging Cell 2012. 11: 968–977. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Becklund BR, Purton JF, Ramsey C, Favre S, Vogt TK, Martin CE, Spasova DS, Sarkisyan G, LeRoy E, Tan JT, Wahlus H, Bondi-Boyd B, Luther SA and Surh CD, The aged lymphoid tissue environment fails to support naive T cell homeostasis. Sci Rep 2016. 6: 30842. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Snyder ME, Finlayson MO, Connors TJ, Dogra P, Senda T, Bush E, Carpenter D, Marboe C, Benvenuto L, Shah L, Robbins H, Hook JL, Sykes M, D’Ovidio F, Bacchetta M, Sonett JR, Lederer DJ, Arcasoy S, Sims PA and Farber DL, Generation and persistence of human tissue-resident memory T cells in lung transplantation. Sci Immunol 2019. 4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Agius E, Lacy KE, Vukmanovic-Stejic M, Jagger AL, Papageorgiou AP, Hall S, Reed JR, Curnow SJ, Fuentes-Duculan J, Buckley CD, Salmon M, Taams LS, Krueger J, Greenwood J, Klein N, Rustin MH and Akbar AN, Decreased TNF-alpha synthesis by macrophages restricts cutaneous immunosurveillance by memory CD4+T cells during aging. J Exp Med 2009. 206: 1929–1940. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Vukmanovic-Stejic M, Chambers ES, Suarez-Farinas M, Sandhu D, Fuentes-Duculan J, Patel N, Agius E, Lacy KE, Turner CT, Larbi A, Birault V, Noursadeghi M, Mabbott NA, Rustin MHA, Krueger JG and Akbar AN, Enhancement of cutaneous immunity during aging by blocking p38 mitogen-activated protein (MAP) kinase-induced inflammation. J Allergy Clin Immunol 2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Brubaker AL, Rendon JL, Ramirez L, Choudhry MA and Kovacs EJ, Reduced neutrophil chemotaxis and infiltration contributes to delayed resolution of cutaneous wound infection with advanced age. J Immunol 2013. 190: 1746–1757. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Rubenfeld GD, Caldwell E, Peabody E, Weaver J, Martin DP, Neff M, Stern EJ and Hudson LD, Incidence and outcomes of acute lung injury. N Engl J Med 2005. 353: 1685–1693. [DOI] [PubMed] [Google Scholar]
  • 58.Thompson WW, Shay DK, Weintraub E, Brammer L, Bridges CB, Cox NJ and Fukuda K, Influenza-associated hospitalizations in the United States. JAMA 2004. 292: 1333–1340. [DOI] [PubMed] [Google Scholar]
  • 59.Cho SJ, Plataki M, Mitzel D, Lowry G, Rooney K and Stout-Delgado H, Decreased NLRP3 inflammasome expression in aged lung may contribute to increased susceptibility to secondary Streptococcus pneumoniae infection. Exp Gerontol 2018. 105: 40–46. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Svedberg FR, Brown SL, Krauss MZ, Campbell L, Sharpe C, Clausen M, Howell GJ, Clark H, Madsen J, Evans CM, Sutherland TE, Ivens AC, Thornton DJ, Grencis RK, Hussell T, Cunoosamy DM, Cook PC and MacDonald AS, The lung environment controls alveolar macrophage metabolism and responsiveness in type 2 inflammation. Nat Immunol 2019. 20: 571–580. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Frankel JH, Boe DM, Albright JM, O’Halloran EB, Carter SR, Davis CS, Ramirez L, Burnham EL, Gamelli RL, Afshar M and Kovacs EJ, Age-related immune responses after burn and inhalation injury are associated with altered clinical outcomes. Exp Gerontol 2018. 105: 78–86. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Brandes M, Klauschen F, Kuchen S and Germain RN, A systems analysis identifies a feedforward inflammatory circuit leading to lethal influenza infection. Cell 2013. 154: 197–212. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Mauad T, Hajjar LA, Callegari GD, da Silva LF, Schout D, Galas FR, Alves VA, Malheiros DM, Auler JO Jr., Ferreira AF, Borsato MR, Bezerra SM, Gutierrez PS, Caldini ET, Pasqualucci CA, Dolhnikoff M and Saldiva PH, Lung pathology in fatal novel human influenza A (H1N1) infection. Am J Respir Crit Care Med 2010. 181: 72–79. [DOI] [PubMed] [Google Scholar]
  • 64.Shivshankar P, Boyd AR, Le Saux CJ, Yeh IT and Orihuela CJ, Cellular senescence increases expression of bacterial ligands in the lungs and is positively correlated with increased susceptibility to pneumococcal pneumonia. Aging Cell 2011. 10: 798–806. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Hinojosa CA, Mgbemena V, Van Roekel S, Austad SN, Miller RA, Bose S and Orihuela CJ, Enteric-delivered rapamycin enhances resistance of aged mice to pneumococcal pneumonia through reduced cellular senescence. Exp Gerontol 2012. 47: 958–965. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Kling KM, Lopez-Rodriguez E, Pfarrer C, Muhlfeld C and Brandenberger C, Aging exacerbates acute lung injury-induced changes of the air-blood barrier, lung function, and inflammation in the mouse. Am J Physiol Lung Cell Mol Physiol 2017. 312: L1–L12. [DOI] [PubMed] [Google Scholar]
  • 67.Bajenoff M, Egen JG, Qi H, Huang AY, Castellino F and Germain RN, Highways, byways and breadcrumbs: directing lymphocyte traffic in the lymph node. Trends Immunol 2007. 28: 346–352. [DOI] [PubMed] [Google Scholar]
  • 68.Luscieti P, Hubschmid T, Cottier H, Hess MW and Sobin LH, Human lymph node morphology as a function of age and site. J Clin Pathol 1980. 33: 454–461. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Thompson HL, Smithey MJ, Uhrlaub JL, Jeftic I, Jergovic M, White SE, Currier N, Lang AM, Okoye A, Park B, Picker LJ, Surh CD and Nikolich-Zugich J, Lymph nodes as barriers to T-cell rejuvenation in aging mice and nonhuman primates. Aging Cell 2019. 18: e12865. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Richner JM, Gmyrek GB, Govero J, Tu Y, van der Windt GJ, Metcalf TU, Haddad EK, Textor J, Miller MJ and Diamond MS, Age-Dependent Cell Trafficking Defects in Draining Lymph Nodes Impair Adaptive Immunity and Control of West Nile Virus Infection. PLoS Pathog 2015. 11: e1005027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Dixit VD, Thymic fatness and approaches to enhance thymopoietic fitness in aging. Curr Opin Immunol 2010. 22: 521–528. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Mau T and Yung R, Adipose tissue inflammation in aging. Exp Gerontol 2018. 105: 27–31. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Lumeng CN, Liu J, Geletka L, Delaney C, Delproposto J, Desai A, Oatmen K, Martinez-Santibanez G, Julius A, Garg S and Yung RL, Aging is associated with an increase in T cells and inflammatory macrophages in visceral adipose tissue. J Immunol 2011. 187: 6208–6216. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Frasca D, Diaz A, Romero M, Vazquez T and Blomberg BB, Obesity induces pro-inflammatory B cells and impairs B cell function in old mice. Mech Ageing Dev 2017. 162: 91–99. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Carter S, Miard S, Caron A, Salle-Lefort S, St-Pierre P, Anhe FF, Lavoie-Charland E, Blais-Lecours P, Drolet MC, Lefebvre JS, Lacombe J, Deshaies Y, Couet J, Laplante M, Ferron M, Bosse Y, Marette A, Richard D, Marsolais D and Picard F, Loss of OcaB Prevents Age-Induced Fat Accretion and Insulin Resistance by Altering B-Lymphocyte Transition and Promoting Energy Expenditure. Diabetes 2018. 67: 1285–1296. [DOI] [PubMed] [Google Scholar]
  • 76.Youm YH, Grant RW, McCabe LR, Albarado DC, Nguyen KY, Ravussin A, Pistell P, Newman S, Carter R, Laque A, Munzberg H, Rosen CJ, Ingram DK, Salbaum JM and Dixit VD, Canonical Nlrp3 inflammasome links systemic low-grade inflammation to functional decline in aging. Cell Metab 2013. 18: 519–532. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Youm YH, Kanneganti TD, Vandanmagsar B, Zhu X, Ravussin A, Adijiang A, Owen JS, Thomas MJ, Francis J, Parks JS and Dixit VD, The Nlrp3 inflammasome promotes age-related thymic demise and immunosenescence. Cell Rep 2012. 1: 56–68. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.McBride MJ, Foley KP, D’Souza DM, Li YE, Lau TC, Hawke TJ and Schertzer JD, The NLRP3 inflammasome contributes to sarcopenia and lower muscle glycolytic potential in old mice. Am J Physiol Endocrinol Metab 2017: ajpendo 00060 02017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Stout-Delgado HW, Cho SJ, Chu SG, Mitzel DN, Villalba J, El-Chemaly S, Ryter SW, Choi AM and Rosas IO, Age-Dependent Susceptibility to Pulmonary Fibrosis Is Associated with NLRP3 Inflammasome Activation. Am J Respir Cell Mol Biol 2016. 55: 252–263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Thakar J, Mohanty S, West AP, Joshi SR, Ueda I, Wilson J, Meng H, Blevins TP, Tsang S, Trentalange M, Siconolfi B, Park K, Gill TM, Belshe RB, Kaech SM, Shadel GS, Kleinstein SH and Shaw AC, Aging-dependent alterations in gene expression and a mitochondrial signature of responsiveness to human influenza vaccination. Aging (Albany NY) 2015. 7: 38–52. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Metcalf TU, Wilkinson PA, Cameron MJ, Ghneim K, Chiang C, Wertheimer AM, Hiscott JB, Nikolich-Zugich J and Haddad EK, Human Monocyte Subsets Are Transcriptionally and Functionally Altered in Aging in Response to Pattern Recognition Receptor Agonists. J Immunol 2017. 199: 1405–1417. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Team H-CSP and Consortium H-I, Multicohort analysis reveals baseline transcriptional predictors of influenza vaccination responses. Sci Immunol 2017. 2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Beerman I, Bock C, Garrison BS, Smith ZD, Gu H, Meissner A and Rossi DJ, Proliferation-dependent alterations of the DNA methylation landscape underlie hematopoietic stem cell aging. Cell Stem Cell 2013. 12: 413–425. [DOI] [PubMed] [Google Scholar]
  • 84.Lemke G and Rothlin CV, Immunobiology of the TAM receptors. Nat Rev Immunol 2008. 8: 327–336. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Rothlin CV, Carrera-Silva EA, Bosurgi L and Ghosh S, TAM receptor signaling in immune homeostasis. Annu Rev Immunol 2015. 33: 355–391. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Qin B, Wang J, Ma N, Yang M, Fu H, Liang Y, Huang F, Yang Z and Zhong R, The association of Tyro3/Axl/Mer signaling with inflammatory response, disease activity in patients with primary Sjogren’s syndrome. Joint Bone Spine 2015. 82: 258–263. [DOI] [PubMed] [Google Scholar]
  • 87.Zheng S, Hedl M and Abraham C, TAM receptor-dependent regulation of SOCS3 and MAPKs contributes to proinflammatory cytokine downregulation following chronic NOD2 stimulation of human macrophages. J Immunol 2015. 194: 1928–1937. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Wang X, Malawista A, Qian F, Ramsey C, Allore HG and Montgomery RR, Age-related changes in expression and signaling of TAM receptor inflammatory regulators in monocytes. Oncotarget 2018. 9: 9572–9580. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Scott RS, McMahon EJ, Pop SM, Reap EA, Caricchio R, Cohen PL, Earp HS and Matsushima GK, Phagocytosis and clearance of apoptotic cells is mediated by MER. Nature 2001. 411: 207–211. [DOI] [PubMed] [Google Scholar]
  • 90.Seitz HM, Camenisch TD, Lemke G, Earp HS and Matsushima GK, Macrophages and dendritic cells use different Axl/Mertk/Tyro3 receptors in clearance of apoptotic cells. J Immunol 2007. 178: 5635–5642. [DOI] [PubMed] [Google Scholar]
  • 91.Wong CK, Smith CA, Sakamoto K, Kaminski N, Koff JL and Goldstein DR, Aging Impairs Alveolar Macrophage Phagocytosis and Increases Influenza-Induced Mortality in Mice. J Immunol 2017. 199: 1060–1068. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Cuervo AM and Wong E, Chaperone-mediated autophagy: roles in disease and aging. Cell Res 2014. 24: 92–104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Metcalf TU, Cubas RA, Ghneim K, Cartwright MJ, Grevenynghe JV, Richner JM, Olagnier DP, Wilkinson PA, Cameron MJ, Park BS, Hiscott JB, Diamond MS, Wertheimer AM, Nikolich-Zugich J and Haddad EK, Global analyses revealed age-related alterations in innate immune responses after stimulation of pathogen recognition receptors. Aging Cell 2015. 14: 421–432. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Rizza W, Veronese N and Fontana L, What are the roles of calorie restriction and diet quality in promoting healthy longevity? Ageing Res Rev 2014. 13: 38–45. [DOI] [PubMed] [Google Scholar]
  • 95.Newman JC, Covarrubias AJ, Zhao M, Yu X, Gut P, Ng CP, Huang Y, Haldar S and Verdin E, Ketogenic Diet Reduces Midlife Mortality and Improves Memory in Aging Mice. Cell Metab 2017. 26: 547–557 e548. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Roberts MN, Wallace MA, Tomilov AA, Zhou Z, Marcotte GR, Tran D, Perez G, Gutierrez-Casado E, Koike S, Knotts TA, Imai DM, Griffey SM, Kim K, Hagopian K, Haj FG, Baar K, Cortopassi GA, Ramsey JJ and Lopez-Dominguez JA, A Ketogenic Diet Extends Longevity and Healthspan in Adult Mice. Cell Metab 2017. 26: 539–546 e535. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Wei M, Brandhorst S, Shelehchi M, Mirzaei H, Cheng CW, Budniak J, Groshen S, Mack WJ, Guen E, Di Biase S, Cohen P, Morgan TE, Dorff T, Hong K, Michalsen A, Laviano A and Longo VD, Fasting-mimicking diet and markers/risk factors for aging, diabetes, cancer, and cardiovascular disease. Sci Transl Med 2017. 9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Jiang W, Zhu Z and Thompson HJ, Dietary energy restriction modulates the activity of AMP-activated protein kinase, Akt, and mammalian target of rapamycin in mammary carcinomas, mammary gland, and liver. Cancer Res 2008. 68: 5492–5499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Sengupta S, Peterson TR, Laplante M, Oh S and Sabatini DM, mTORC1 controls fasting-induced ketogenesis and its modulation by ageing. Nature 2010. 468: 1100–1104. [DOI] [PubMed] [Google Scholar]
  • 100.Darcy J, McFadden S and Bartke A, Altered structure and function of adipose tissue in long-lived mice with growth hormone-related mutations. Adipocyte 2017: 1–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Riera CE and Dillin A, Tipping the metabolic scales towards increased longevity in mammals. Nat Cell Biol 2015. 17: 196–203. [DOI] [PubMed] [Google Scholar]
  • 102.Youm YH, Nguyen KY, Grant RW, Goldberg EL, Bodogai M, Kim D, D’Agostino D, Planavsky N, Lupfer C, Kanneganti TD, Kang S, Horvath TL, Fahmy TM, Crawford PA, Biragyn A, Alnemri E and Dixit VD, The ketone metabolite beta-hydroxybutyrate blocks NLRP3 inflammasome-mediated inflammatory disease. Nat Med 2015. 21: 263–269. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Edwards C, Canfield J, Copes N, Rehan M, Lipps D and Bradshaw PC, D-beta-hydroxybutyrate extends lifespan in C. elegans. Aging (Albany NY) 2014. 6: 621–644. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Pearce EL and Pearce EJ, Metabolic pathways in immune cell activation and quiescence. Immunity 2013. 38: 633–643. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.O’Neill LA and Pearce EJ, Immunometabolism governs dendritic cell and macrophage function. J Exp Med 2016. 213: 15–23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Camell C, Goldberg E and Dixit VD, Regulation of Nlrp3 inflammasome by dietary metabolites. Semin Immunol 2015. 27: 334–342. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Serhan CN, Pro-resolving lipid mediators are leads for resolution physiology. Nature 2014. 510: 92–101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Arnardottir HH, Dalli J, Colas RA, Shinohara M and Serhan CN, Aging delays resolution of acute inflammation in mice: reprogramming the host response with novel nano-proresolving medicines. J Immunol 2014. 193: 4235–4244. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Cai B, Kasikara C, Doran AC, Ramakrishnan R, Birge RB and Tabas I, MerTK signaling in macrophages promotes the synthesis of inflammation resolution mediators by suppressing CaMKII activity. Sci Signal 2018. 11. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES