Next Article in Journal
Modeling of Protein Structural Flexibility and Large-Scale Dynamics: Coarse-Grained Simulations and Elastic Network Models
Next Article in Special Issue
Structure and Physiological Regulation of AMPK
Previous Article in Journal
Speciation of Selenium in Brown Rice Fertilized with Selenite and Effects of Selenium Fertilization on Rice Proteins
Previous Article in Special Issue
AMP-Activated Protein Kinase (AMPK)-Dependent Regulation of Renal Transport
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

AMP-Activated Protein Kinase and Host Defense against Infection

1
Department of Microbiology, Chungnam National University School of Medicine, Daejeon 35015, Korea
2
Infection Control Convergence Research Center, Chungnam National University School of Medicine, Daejeon 35015, Korea
3
Department of Medical Science, Chungnam National University School of Medicine, Daejeon 35015, Korea
4
Department of Infection Biology, Chungnam National University School of Medicine, Daejeon 35015, Korea
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2018, 19(11), 3495; https://doi.org/10.3390/ijms19113495
Submission received: 27 September 2018 / Revised: 5 November 2018 / Accepted: 5 November 2018 / Published: 6 November 2018
(This article belongs to the Special Issue AMP-Activated Protein Kinase Signalling)

Abstract

:
5′-AMP-activated protein kinase (AMPK) plays diverse roles in various physiological and pathological conditions. AMPK is involved in energy metabolism, which is perturbed by infectious stimuli. Indeed, various pathogens modulate AMPK activity, which affects host defenses against infection. In some viral infections, including hepatitis B and C viral infections, AMPK activation is beneficial, but in others such as dengue virus, Ebola virus, and human cytomegaloviral infections, AMPK plays a detrimental role. AMPK-targeting agents or small molecules enhance the antiviral response and contribute to the control of microbial and parasitic infections. In addition, this review focuses on the double-edged role of AMPK in innate and adaptive immune responses to infection. Understanding how AMPK regulates host defenses will enable development of more effective host-directed therapeutic strategies against infectious diseases.

1. Introduction

5′-AMP-activated protein kinase (AMPK) is an intracellular serine/threonine kinase and a key energy sensor that is activated under conditions of metabolic stress [1,2,3]. It governs a variety of biological processes for the maintenance of energy homeostasis in response to metabolic stresses such as adenosine triphosphate (ATP) depletion [2]. Due to its critical function in metabolic homeostasis, much research has focused on the roles of AMPK in metabolic diseases and cancers [4,5,6]. However, much less is known about the function of AMPK in infection [7]. Due to the energetic demands of infected cells, most infections by intracellular pathogens are associated with activation of host AMPK, presumably to promote microbial proliferation [8]. AMPK functions as a modulator of host defenses against intracellular bacterial, viral, and parasitic infections [9,10,11,12]. Indeed, numerous viruses have the ability to trigger metabolic changes, thereby modulating AMPK activity and substrate selection [13], and AMPK signaling could facilitate or inhibit intracellular viral replication depending on the virus infection [14].
This review focuses on the double-edged role of AMPK in the regulation of host antimicrobial defenses in infections of viruses, bacteria, and parasites. In this review, we describe the existing evidence for the defensive and inhibitory roles of AMPK and the mechanisms underlying its regulation of innate and inflammatory responses. Finally, we describe AMPK-targeting agents that enhance host defenses against infection or control harmful inflammation.

2. Overview of AMPK

5′-AMP-activated protein kinase (AMPK), a serine/threonine kinase, is a key player in bioenergetic homeostasis to preserve cellular ATP [1]. AMPK is activated in response to an increased cellular adenosine monophosphate (AMP)/ATP or adenosine diphosphate (ADP)/ATP ratio, thus promoting catabolic pathways and suppressing biosynthetic pathways [1,2,3]. Mammalian AMPK exists as a heterotrimeric complex comprising a catalytic subunit α (α1 and α2), a scaffolding β subunit (β1 and β2), and a regulatory γ subunit (γ1, γ2, and γ3) (Figure 1A) [15]. Multiple isoforms of AMPK are encoded by distinct genes of the subunit isotypes, depending on the cell/tissue or species [16]. The AMPK subunit composition and ligand-induced activities of each AMPK isoform complex can differ among cell types, although the α1, β1, and γ1 isoforms are ubiquitously expressed [16,17].
The AMPK α subunit contains a kinase domain at the N terminus, which is activated by phosphorylation of Thr-172 by the major upstream liver kinase B1 (LKB1) [18,19]. In contrast to LKB1, the upstream Ca2+-calmodulin-dependent kinase kinase (CaMKK) activates AMPK in response to an increased intracellular Ca2+ concentration in the absence of significant changes in ATP/ADP/AMP levels [20]. The regulatory β subunit of AMPK contains a glycogen-binding domain that can sense the structural state of glycogen [21]. Four consecutive cystathionine-β-synthase domains in the regulatory γ subunit are essential for binding to adenosine nucleotides to form an active αβγ complex (Figure 1B) [22,23].
Different AMPK isoforms may have distinct biological functions in different physiological and pathological systems. AMPK governs the cellular energy status by acting as a crucial regulator of energy homeostasis in response to various metabolic stresses, including starvation, hypoxia, and muscle contraction. AMPK activity can be altered by numerous factors, including hormones, cytokines, and nutrients, as well as diverse pathological changes such as metabolic disturbances [24,25]. Because AMPK is important in the adaptation to energy stress, dysregulation of or decreased AMPK activation is implicated in the development of metabolic disorders associated with insulin resistance [6]. In addition to its primary role in the regulation of energy metabolism, AMPK signaling plays a critical role in host–microbial interactions [7]. Furthermore, infections by several viruses result in dysregulation or stimulation of AMPK activity [13]. In mycobacterial infections, AMPK activation promotes activation of host defenses in macrophages and in vivo [12,26]. However, much less is known about the function of AMPK in innate host defenses compared with that in the regulation of metabolism and its mitochondrial function.

3. Multifaceted Role of AMPK in Antimicrobial Responses

Viruses have evolved strategies to manipulate the AMPK signaling pathway to escape host defenses. Indeed, several pathogens can modulate the activity of AMPK/mTOR to obtain sufficient energy for their growth and proliferation [8]. In this review, we discuss microbial manipulation of AMPK activity to affect host defenses against infections. Figure 2 summarizes the multiple roles of AMPK in the viral and bacterial infections addressed in this review. The detailed mechanisms and outcomes of host–pathogen interactions in terms of AMPK modulation are described in Table 1, Table 2, Table 3 and Table 4.

3.1. Roles of AMPK in Viral Infections

3.1.1. Beneficial Effects of AMPK on Virus Infections

Hepatitis C virus (HCV) is a major etiologic agent of chronic liver disease worldwide. HCV infection inhibits AMPKα phosphorylation and signaling [27], and the AMPK agonist metformin suppresses HCV replication in an autophagy-independent manner [28]. Moreover, HCV core protein increases the levels of reactive oxygen species (ROS) and alters the NAD/NADH ratio to decrease the activity and expression of sirtuin 1 (SIRT1) and AMPK, thereby altering the metabolic profile of hepatocytes. This mechanism is implicated in the pathogenesis of hepatic metabolic diseases [29]. As AMPK is a crucial regulator of lipid and glucose metabolism, pharmacological restoration of AMPK activity inhibits lipid accumulation and viral replication in HCV-infected cells [27]. In addition, metformin enhances type I interferon (IFN) signaling by activating AMPK, resulting in inhibition of HCV replication [30]. AMPK inhibition resulted in the downregulation of type I IFN signaling and rescue of the metformin-mediated decrease in the HCV core protein level [30]. Moreover, the AMPK activator 5-aminoimidazole-4-carboxamide 1-β-d-ribofuranoside (AICAR) inhibits HCV replication by activating AMPK signaling, although the anti-HCV effect of metformin is independent of AMPK activation [31]. In chronic HCV infection, the expression of Sucrose-non-fermenting protein kinase 1/AMP-activated protein kinase-related protein kinase (SNARK), an AMPK-related kinase, is increased to promote transforming growth factor β signaling, which is critical for hepatic fibrogenesis [32]. A more recent study showed that HCV-mediated ROS production triggers AMPK activation to attenuate lipid synthesis and promote fatty acid β-oxidation in HCV-infected cells [33]. These data suggest that HCV inhibits AMPK activation to promote its replication, and that the restoration of AMPK activity may be an effective therapeutic modality for HCV infection that acts by metabolic reprogramming or modulation of type I IFN production in host cells [27,28,30,33].
In hepatitis B viral (HBV) infection, AMPK can promote or inhibit viral replication. The detrimental effects of AMPK are described in the following section. Xie et al. reported that AMPK, which is activated by HBV-induced ROS accumulation, suppresses HBV replication [9]. Mechanistically, AMPK activation leads to HBV-mediated autophagic activation, which enhances autolysosome-dependent degradation to restrict viral proliferation [9]. AMPK activity is also involved in the defense against vesicular stomatitis virus, the causal agent of an influenza-like illness, by activating stimulator of IFN genes (STING) [10]. Treatment of mouse macrophages or fibroblasts with an AMPK inhibitor suppressed the production of type I IFN and TNF-α in response to a STING-dependent ligand or agonist, suggesting a role for AMPK in STING signaling [10]. AMPK plays a role in the excessive inflammatory cytokine/chemokine levels in Mint3/Apba3 depletion models of severe pneumonia due to influenza virus [34]. Indeed, food-derived polyphenols, such as epigallocatechin gallate and curcumin, are useful for controlling viral and bacterial infections [35]. Although a review of AMPK-modulating polyphenols is beyond our scope, we highlight the therapeutic promise of polyphenols against infection. For example, curcumin from Curcuma longa inhibits influenza A viral infection in vitro and in vivo, at least in part by activating AMPK [36]. The polyphenol epigallocatechin gallate attenuates Tat-induced human immunodeficiency virus (HIV)-1 transactivation by activating AMPK [37]. Further studies should examine the ability of food-derived polyphenols to activate AMPK signaling to control viral replication in host cells.
Human adenovirus type 36, which is associated with obesity, inhibits fatty acid oxidation and AMPK activity and increases accumulation of lipid droplets in infected cells [38]. The AMPK signaling pathway and its upstream regulator LKB1 repress replication of the bunyavirus Rift Valley Fever virus (RVFV), a re-emerging human pathogen [39]. The mechanisms of the antiviral effects of AMPK on RVFV and other viruses are mediated by AMPK inhibition of fatty acid synthesis [39]. Pharmacologic activation of AMPK suppresses RVFV infection and reduces lipid levels by inhibiting fatty acid biosynthesis [39]. In addition, the AMPK/Sirt1 activators resveratrol and quercetin significantly reduce the viral titer and gene expression, as well as increase the viability of infected neurons, in herpes simplex virus type 1 (HSV-1) infection [40]. Moreover, coxsackievirus B3 (CVB3) infection triggers AMPK activation, which suppresses viral replication in HeLa and primary myocardial cells [41]. The AMPK agonists AICAR and metformin suppress CVB3 replication and attenuate lipid accumulation by inhibiting lipid biosynthesis [41]. Thus, regulation of fatty acid metabolism by AMPK signaling is an essential component of cell autonomous immune responses [39].
Latent membrane protein 1 (LMP1) of Epstein-Barr virus (EBV) inactivates LKB1/AMPK, whereas AMPK activation by AICAR abrogated LMP1-mediated proliferation and transformation of nasopharyngeal epithelial cells, suggesting therapeutic potential for EBV-associated nasopharyngeal carcinoma [42]. Moreover, constitutive activation of AMPK inhibited lytic replication of Kaposi’s sarcoma-associated herpesvirus in primary human umbilical vein endothelial cells [43]. These data suggest that AMPK suppresses cell transformation and infection-related tumorigenesis in a context-dependent manner. The roles of AMPK in viral infection are listed in Table 1.

3.1.2. Detrimental Effects of AMPK on Virus Infections

Several viruses manipulate AMPK signaling to promote their replication. Genome-scale RNA interference screening of host factors in rotaviral infection identified AMPK as a critical factor in the initiation of a rotavirus-favorable environment [44]. In dengue viral infections, the 3-hydroxy-3-methylglutaryl-CoA reductase (HMGGR) activity elevated by AMPK inactivation resulted in generation of a cholesterol-rich environment in the endoplasmic reticulum, which promoted formation of viral replication complexes [45]. Also, dengue viral infection stimulates AMPK activation to induce proviral lipophagy, thereby enhancing fatty acid β-oxidation and viral replication [46].
In HBV infection, the HBV X protein activates AMPK, and inhibition of AMPK reduces HBV replication in rat primary hepatocytes [47]. Inhibition of AMPK led to activation of mTORC1, which is required for inhibition of HBV replication in the presence of low AMPK activity [47]. The crosstalk between AMPK and mTORC1 may enable development of therapeutics that suppress HBV replication and thus also hepatocellular carcinoma (HCC) development [47]. However, as described in Section 3.1.1, AMPK activation by AICAR inhibits extracellular HBV production in HepG2 cells [9]. The discrepancy may be attributed to the use of different cell lines in the two studies [9,47]. Further work should address the role of AMPK in HBV infection in vitro and in vivo. In infection by Zaire Ebolavirus (EBOV), the expression levels of the γ2 subunit of AMPK are correlated with EBOV transduction in host cells. In mouse embryonic fibroblasts treated with a small-molecule inhibitor of AMPK (compound C), it was shown that AMPK activity is required for EBOV replication in host cells and EBOV glycoprotein-mediated entry/uptake [48]. In addition, Avian reoviral infection upregulates AMPK phosphorylation, which leads to activation of mitogen-activated protein kinase (MAPK) p38 in Vero cells, which enhances viral replication [49]. The nonstructural protein p17 of avian reovirus positively regulates AMPK activity, inducing autophagy and increasing viral replication [50].
In HSV-1 infection, the activated AMPK/Sirt1 axis inhibits host-cell apoptosis during early-stage infection, which promotes viral latency and protects neurons [51]. However, during the later stages of infection, HSV-1 induces apoptosis of host cells concomitantly with Sirt1 activation [51]. In HIV-infected cocaine abusers, AMPK signaling plays a role in energy deficit and neuronal dysfunction, which are associated with the development of neuroAIDS [52]. These data suggest that differential regulation of AMPK signaling is a determinant of the viral infection course.
Using a kinome-profiling approach, AMPK and related kinases were found to be effectors of human cytomegalovirus (HCMV) replication [53,54]. HCMV infection induces AMPK and CaMKK2 (upstream activator of AMPK)-dependent remodeling of core metabolism, both of which are required for optimal yield and replication of HCMV [53,54]. Notably, inhibition of AMPK activity by short-interfering RNA-mediated AMPK knockdown or an AMPK antagonist (compound C) prevents viral gene expression, providing valuable insight into the mechanisms of HCMV infection [53]. In addition, the AMPK activation-dependent modulation by HCMV of host-cell metabolism is associated with HCMV replication [55]. HCMV-mediated AMPK activation is dependent on CaMKK, and inhibition of AMPK activity abrogated HCMV replication and DNA synthesis [55]. Furthermore, the cardiac glycoside digitoxin induces phosphorylation of AMPK/ULK1, whereas it suppresses mTOR activity to increase autophagic flux and inhibit HCMV replication [56]. Moreover, HCMV induces production of the host protein viperin [57], which is required for AMPK activation, transcriptional activation of GLUT4 and lipogenic enzymes, and lipid synthesis [58]. The enhanced lipid synthesis promotes formation of the viral envelope and production of HCMV virions [58].
In infection, host-cell autophagy plays an important role in host defense and virus survival. Several viruses can manipulate or subvert the autophagic machinery to favor viral replication. For example, respiratory syncytial virus activates autophagy via the AMPK/mTOR signaling pathway to enhance its replication by inhibiting host-cell apoptosis [59]. Bluetongue virus, a double-stranded segmented RNA virus, also induces host-cell autophagy by activating AMPK [60]. Moreover, AMPK is an upstream regulator of rabies virus-induced incomplete autophagy to provide the scaffolds for viral replication [61]. In Sendai viral infection, AMPK activity is required for autophagic initiation to promote viral replication [62]. In oncogenic EBV infection, the increased cell survival caused by AMPK-mediated autophagic activation maintains early hyperproliferation of infected cells [63]. These data suggest AMPK activity to be a therapeutic target for the development of novel antiviral agents.
Importantly, type I IFN, a critical effector in the antiviral response, attenuates AMPK phosphorylation and increases the intracellular ATP level [64]. In addition, IFN-β-mediated glycolytic metabolism is important for the acute phase of the antiviral response to CVB3 [64]. The antiviral cytokine IFN-β regulates host-cell metabolism to enhance glucose uptake and ATP generation, which promote the antiviral response [64]. In infection by snakehead vesiculovirus, miR-214 targeting AMPK suppressed viral replication and upregulated IFN-α expression [65]. Thus, regulation of AMPK activity by the host–pathogen interaction mediates diverse metabolic effects, which modulate viral replication and the host defense response. The beneficial and detrimental effects of AMPK on viral infections are summarized in Table 1 and Table 2, respectively.

3.2. Bacterial Infections and AMPK Activation

Intracellular pathogens manipulate the AMPK signaling pathway to alter their metabolic environment to favor bacterial survival or pathogenesis. Mitochondrial dysfunction triggers AMPK signaling, thus enhancing the proliferation of Legionella pneumophila, a respiratory pathogen, in Dictyostelium cells [66]. Inhibition of AMPK activation reversed the increased Legionella proliferation in host cells with mitochondrial disease [66]. However, the AMPK activator metformin triggers mitochondrial ROS generation and activates the AMPK signaling pathway to enhance the host response to L. pneumophila in macrophages and promote survival in a murine model of L. pneumophila pneumonia [67]. Thus, the role of AMPK activation in bacterial infections differs depending on the host species.
Salmonella typhimurium degrades SIRT1/AMPK to evade host xenophagy [68]. In addition, cytosolic Salmonella is ubiquitinated and targeted for xenophagy by AMPK activation [69,70]. AMPK activation by AICAR induces autophagy and colocalization of Salmonella-containing vacuoles with LC3 autophagosomes [68], whereas inhibition of AMPK by compound C increases bacterial replication by suppressing autophagy [69]. In Salmonella-infected cells, AMPK activation is mediated through toll-like receptor-activated TGF-β-activated kinase 1 (TAK1) [69] which is a direct upstream kinase of AMPK in addition to LKB1 and CaMKK2 [71]. In Brucella abortus infection, AMPK activation enhances intracellular growth of B. abortus by inhibiting nicotinamide adenine dinucleotide phosphate (NADPH) oxidase-mediated ROS generation [72]. In models of Escherichia coli sepsis, ATP-induced pyroptosis is blocked by piperine, a phytochemical present in black pepper (Piper nigrum Linn) [73]. The inhibitory effects of piperine on pyroptosis and systemic inflammation are mediated by regulation of the AMPK signaling pathway, as shown by suppression of ATP-mediated AMPK activation by piperine treatment in vitro and in vivo [73]. These data suggest that AMPK plays multiple roles in bacterial infections.
AMPK activation promotes host defenses against infections by several microbes. Transcriptomic and proteomic analyses of a Caenorhabditis elegans model indicated that AMPKs function as regulators and mediators of the immune response to infection by, for example, Bacillus thuringiensis [74]. Several small-molecule AMPK activators exert protective effects against Helicobacter pylori-induced apoptosis of gastric epithelial cells [70,75]. The AMPK agonists A-769662 and resveratrol, as well as AMPKα overexpression, inhibit apoptosis in H. pylori-infected gastric epithelial cells [76]. The AMPK activator compound 13 ameliorates H. pylori-induced apoptosis of gastric epithelial cells by modulating ROS levels via the AMPK-heme oxygenase-1 axis [76]. Blockade of AMPK signaling significantly abrogates the protective effect of compound 13 against H. pylori within gastric epithelial cells [76].
Metformin and AICAR repress infection by Pseudomonas aeruginosa, an important opportunistic pathogen, of airway epithelial cells by inhibiting bacterial growth and increasing transepithelial electrical resistance [77]. AMPKα1 depletion increased the susceptibility to Staphylococcus aureus endophthalmitis in mice [78], suggesting a protective role for AMPK in bacterial retinal inflammation. Moreover, AMPK activation by AICAR enhances its anti-inflammatory effects, phagocytosis, and bactericidal activity against S. aureus infection of various phagocytic cells including microglia, macrophages, and neutrophils [78]. Epigallocatechin gallate, a polyphenol in green tea, inhibits the viability of Propionibacterium acnes, a pathogen associated with acne, and exerted an antilipogenic effect in SEB-1 sebocytes by activating the AMPK/sterol regulatory element-binding protein pathway [35]. Moreover, Acinetobacter baumannii, an emerging opportunistic pathogen, activates autophagy via the AMPK/ERK/mTOR pathway to promote an antimicrobial response to intracellular A. baumannii [79]. The roles of AMPK in bacterial infection are summarized in Table 3.

3.3. Roles of AMPK in Mycobacterial Infection

The seminal study by Singhal et al. addressed the effect of metformin as an adjunctive therapy for tuberculosis. Importantly, metformin suppressed the intracellular growth of Mycobacterium tuberculosis (Mtb) in vitro, including drug-resistant strains, by activating the AMPK signaling pathway [80]. In vivo, metformin attenuated the immunopathology and enhanced the immune response and showed a synergistic effect with conventional anti-TB drugs in Mtb-infected mice [80]. The microbicidal effect of metformin in macrophages is due, at least in part, to mitochondrial ROS generation, which is associated with AMPK signaling [80]. Type 2 diabetes mellitus (DM) is re-emerging as a risk factor for human tuberculosis, thus candidate host directed therapeutic targets for tuberculosis combined with DM should be identified [85]. Human cohort studies showed that metformin treatment for DM is associated with a decreased prevalence of latent tuberculosis compared with alternative DM treatments, suggesting metformin to be a candidate HDT for tuberculosis patients with type 2 DM [80,85]. Indeed, Mtb infection inhibits AMPK phosphorylation but increases mTOR kinase activation in macrophages [12]. The AMPK activator AICAR via autophagic activation enhances phagosomal maturation and antimicrobial responses in macrophages in Mtb infection [12]. In human monocytes/macrophages, vitamin D-mediated antimicrobial responses are mediated by the antimicrobial peptide LL-37 via AMPK activation [81]. In addition, LL-37-induced autophagy by phenylbutyrate, alone or in combination with vitamin D, promotes intracellular killing of Mtb in human macrophages via AMPK- and PtdIns3K-dependent pathways [82]. Recent findings revealed the role of gamma-aminobutyric acid (GABA) in AMPK activation to enhance the autophagy and the antimicrobial responses [83]. Silencing of AMPK by a lentiviral short hairpin RNA (shRNA) specific to AMPK reduces GABA-induced autophagic activation as well as phagosomal maturation during Mtb infection [83].
MicroRNAs are small non-coding RNAs involved in the regulation of diverse physiological and pathological processes, including Mtb infection. Mycobacterial infection of macrophages upregulates miR-33 and miR-33*, which target and suppress AMPKα [86]. Interestingly, miR-33/miR-33* regulates autophagy by suppressing AMPK-dependent activation of the transcription of autophagy- and lysosome-related genes and promoting accumulation of lipid bodies in Mtb infection [86]. Mtb infection increases the expression of MIR144*/has-miR-144-5p, which targets DNA damage regulated autophagy modulator 2 (DRAM2), to inhibit the antimicrobial responses to Mtb infection in human monocytes/macrophages. In contrast, autophagic activators enhance production of the autophagy-related protein DRAM2 by activating the AMPK signaling pathway; this contributes to host defenses against Mtb in human macrophages [87].
Although AMPK may play a protective role in tuberculosis, it has also been reported to exert an immunopathological effect by driving the secretion of neutrophil matrix metalloproteinase-8 (MMP-8), resulting in matrix destruction and cavitation, which enhance the spread of Mtb [84]. Neutrophil-derived MMP-8 secretion is upregulated in Mtb infection and neutrophils from AMPK-deficient patients express lower levels of MMP-8, suggesting a key role for MMP-8 in tuberculosis immunopathology [84]. Because the pathogenesis of tuberculosis is complex, further information on the function of AMPK in the immune response to Mtb infection is needed for development of improved therapeutic strategies [88]. The roles of AMPK in mycobacterial infection are listed in Table 3.

3.4. Roles of AMPK in Parasite Infections

The immune response to parasitic helminths involves M2-type cells, CD4(+) Th2 cells, and group 2 innate lymphoid cells. AMPK activation regulates type 2 immune responses and ameliorates lung injury in response to hookworm infections [89]. Mice deficient in AMPK α1 subunit exhibited impaired type 2 responses, an increased intestinal worm burden, and exacerbated lung injury [89]. In Leishmania-infected macrophages, Leishmania infantum causes a metabolic switch to enhance oxidative phosphorylation by activating LKB1/AMPK and SIRT1 [90]. Impairment of metabolic reprogramming by SIRT1 or AMPK suppresses intracellular growth of the parasite, suggesting a role for AMPK/SIRT1 in intracellular proliferation of L. infantum [90]. In Schistosoma japonicum egg antigen (SEA)-mediated autophagy, which is modulated by IL-7 and the AMPK signaling pathway, ameliorate liver pathology, suggesting AMPK to be a therapeutic target factor for schistosomiasis [91].
Notably, host AMPK activity is decreased by hepatic Plasmodium infection. Activation of the AMPK signaling pathway by AMPK agonists, including salicylate, suppresses the intracellular replication of malaria parasites, including that of the human pathogen Plasmodium falciparum [92]. These data suggest that host AMPK signaling is a therapeutic target for hepatic Plasmodium infection [92]. In addition, resveratrol protects cardiac function and reduces lipid peroxidation and trypanosomal burden in the heart by activating AMPK, suggesting a role for AMPK in Chagas heart disease [93]. The roles of AMPK in parasitic infections are listed in Table 4.

3.5. AMPK in Fungal Infection

A recent phosphoproteomic analysis of Cryptococcus neoformans (Cn) infection showed that AMPK activation is triggered by fungal phagocytosis and is required for autophagic induction. Interestingly, AMPK depletion in monocytes promoted host resistance to fungal infection in mouse models, suggesting that AMPK represses the immune response to Cryptococcus infection [94].

4. Roles of AMPK in Innate and Adaptive Immune Responses

The roles of AMPK in modulation of the mitochondrial network and energy metabolism, which are associated with the immune response, have been investigated [95]. Here, we briefly review recent data on AMPK regulation of innate and adaptive immune responses in infection and inflammation. Figure 3 summarizes the regulatory roles and mechanisms of a variety of small-molecule AMPK activators in terms of innate immune and inflammatory responses.
The data must be interpreted cautiously, as the small-molecule activators (e.g., AICAR, metformin, and compound C) function via off-target mechanisms, such as AMPK-independent pathways or inhibition of protein kinases other than AMPK [31,96,97]. The beneficial effects of these compounds remain to be fully determined. Thus, selective compounds such as MK-8722 [98] and SC4 [99] and the selective inhibitor SBI-0206965 [100] should be considered for future AMPK-targeted treatment strategies.

4.1. Role of AMPK in Regulation of the Innate Immune Response

AMPK is involved in regulation of the innate immune response. For example, the innate-immune stimulator toll-like receptor (TLR) 9 inhibits energy substrates (intracellular ATP levels) and activates AMPK, which enhances stress tolerance in cardiomyocytes and neurons, while stimulation by the TLR9 ligand induces inflammation [101]. The AMPK activator AICAR suppresses the lung inflammation induced by lipoteichoic acid, a major component of the cell wall of Gram-positive bacteria [102]. Natural killer (NK) cells are crucial in the innate immune response to viral infections and transformed cells. Activation of the AMPK signaling pathway or inhibition of mTOR is associated with enhanced mitophagy and an increased number of memory NK cells in antiviral responses [103]. In contrast, increased expression of the inhibitory killer cell lectin-like receptor G1 in aged humans is related to AMPK activation, which has been implicated in disruption of NK cell function [104]. In addition, AMPK activation contributes to CD1d-mediated activation of NK T cells, an important cell type in the innate immune response [105]. The findings above suggest that AMPK plays a pleiotropic role in the regulation of the innate immune response depending on the stimulus and cell type in question.

4.2. AMPK Regulation of Local and Systemic Inflammation

AMPK activators enhance neutrophil chemotaxis, phagocytosis, and bacterial killing to protect against peritonitis-induced sepsis [106]. Indeed, AMPK activators including metformin inhibit injurious inflammatory responses, including neutrophil proinflammatory responses and injury to multiple organs such as the lung, liver, and kidney [107,108,109]. Pharmacologic activation of AMPK by metformin, berberine, or AICAR dampens excessive TLR4/NF-κB signaling, M2-type macrophage polarization, and the production of proinflammatory mediators in vitro and in models of sepsis [110,111,112,113,114,115]. The anti-inflammatory effect of metformin in mice with lipopolysachharide (LPS)-induced septic shock and in ob/ob mice is mediated at least in part by AMPK activation [116]. In septic mice, AMPK activation by AICAR or metformin reduces the severity of sepsis-induced lung injury, enhances AMPK phosphorylation in the brain, and attenuates the inflammatory response [117,118].
Treatment with trimetazidine protects against LPS-induced myocardial dysfunction, exerts an anti-apoptotic effect, and attenuates the inflammatory response due to its effect on the SIRT1/AMPK pathway [119]. Moreover, the flavonoid naringenin dampens inflammation in vitro and protects against murine endotoxemia in vivo; these effects are mediated by AMPK/ATF3-dependent inhibition of the TLR4 signaling pathway [120]. In severe acute HBV infection, halofuginone, a plant alkaloid, inhibits viral replication by activating AMPK-mediated anti-inflammatory responses [121]. AMPK activation also enhances the phagocytic capacities of neutrophils and macrophages [122]. Transient receptor potential melastatin 2, an oxidant sensor cation channel, promotes extracellular trap formation by neutrophils via the AMPK/p38 MAPK pathway, enhancing their antimicrobial activity [123]. AMPK activation not only modulates the acute inflammatory response but also promotes neutrophil-dependent bacterial uptake and killing [106].
In perinatal hypoxic–ischemic encephalopathy, prolonged activation of AMPK signaling suppresses the response to oxygen/glucose deprivation and promotes neonatal hypoxic–ischemic injury [124]. Although AMPK inhibition increases neuronal survival, blockade of AMPK prior to oxygen/glucose deprivation increases cell damage and death [124]. Therefore, the clinical implications of AMPK activation are complex, and further preclinical and clinical data are needed to enable therapeutic use of AMPK activators in patients with acute or chronic inflammation.

4.3. Role of AMPK in Inflammasome Activation

AMPK is implicated in modulation of NLRP3 inflammasome activation. The bactericidal activity of the isoquinoline alkaloid berberine exerts a bactericidal effect by augmenting inflammasome activation via AMPK signaling [125]. However, metformin increases mortality of mice with bacteremia, likely via an AMPK-mediated increase in ATP-induced inflammasome activation and pyroptosis [126].
AMPK is implicated in the inhibition of palmitate-induced inflammasome activation [127]. The AMPK activator AICAR inhibits palmitate-induced activation of the NLRP3 inflammasome and IL-1β secretion by suppressing ROS generation [127]. In addition, NLRP3 inflammasome activation and production of IL-1β are upregulated in the peripheral mononuclear cells of drug-naïve type-2 diabetic patients, suggesting a role of the inflammasome in the pathogenesis of type-2 diabetes [128]. Interestingly, AMPK activation is responsible for the significantly reduced mature IL-1β level in peripheral myeloid cells from type-2 diabetic patients after two months of metformin therapy [128]. In a model of hyperalgesia, which is associated with NLRP3 inflammasome activation, metformin attenuated the clinical symptoms and improved the biochemical parameters, whereas blockade of AMPK activation by compound C provoked hyperalgesia and increased the levels of IL-1β and IL-18 [129]. Furthermore, pharmacological activation of AMPK inhibits the monosodium urate (MSU) crystal-induced inflammatory response, suggesting a role for AMPK in gouty inflammation. Moreover, colchicine, an inhibitor of microtubule assembly used to treat gouty arthritis, enhances AMPKα-mediated phosphorylation, thereby inhibiting inflammasome activation and IL-1β release [130]. Further studies on the efficacy of AMPK activators against inflammasome-associated diseases are thus warranted.

4.4. Role of AMPK in the Regulation of the Adaptive Immune Responses

AMPKα1 is a key regulator of the adaptive immune response, particularly T helper (Th1) 1 and Th17 cell differentiation and the T-cell responses to viral and bacterial infections [131]. In addition, in models of simian immunodeficiency viral infection, AMPK activation is associated with the virus-specific CD8(+) cytotoxic T-lymphocyte population and control of Simian Immunodeficiency Virus (SIV) [132]. The mechanism(s) by which AMPK signaling activates innate and adaptive immune responses and controls excessive inflammation must be determined if the potential of AMPK-targeted therapy is to be realized.

5. Conclusions

Although much research has focused on the role of AMPK in the regulation of mitochondrial and metabolic homeostasis, several issues remain to be addressed. Further work should focus on the mechanism(s) by which AMPK modulates host defenses against infections in vivo. Several pathogens modulate the host metabolic environment to promote their survival and replication. Because of its role in regulating mitochondrial metabolism, dynamics, and biogenesis, AMPK signaling can provide energy to the pathogen and/or host, benefitting either. Stimulation of AMPK activity enhances host defenses against diverse viruses, bacteria, and parasites, notably Mtb. Moreover, AMPK links the innate and adaptive immune responses to infection. However, the molecular mechanisms underlying AMPK regulation of innate and adaptive immunity are unclear. AMPK-targeted small molecules have potential as antimicrobial agents as well as metabolic drugs. Further work is needed to enable development of therapeutics that target AMPK to control inflammation and promote host defenses against infection. This work should focus on elucidating the mechanisms by which AMPK and/or AMPK-targeting compounds modulate host defenses against infection.

Funding

We are indebted to current and past members of our laboratory for discussions and investigations that contributed to this article. We apologize to colleagues whose work and publications could not be referenced owing to space constraints. This work was supported by the National Research Foundation of Korea (NRF) Grant funded by the Korean Government (MSIP) (No. 2017R1A5A2015385) and by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIP) (No. NRF-2015M3C9A2054326).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ADPAdenosine diphosphate
AICAR5-aminoimidazole-4-carboxamide 1-β-d-ribofuanoside
AMPAdenosine monophosphate
AMPK5′-AMP-activated protein kinase
ATF3Activating transcription factor 3
ATPAdenosine triphosphate
CaMKKCa2+/calmodulin-dependent protein kinase kinases
CCL17Chemokine ligand 17
CVB3Coxsackie virus B3
DMDiabetes mellitus
EBOVZaire Ebolavirus
EBVEpstein-Barr Virus
GABAGamma-aminobutyric acid
HBVHepatitis B virus
HCCHepatocellular carcinoma
HCMVHuman cytomegalovirus
HCVHepatitis C virus
HIVHuman immunodeficiency virus
HMGCR3-hydroxy-3-methylglutaryl-CoA reductase
HSV-1Herpex Simplex Virus Type 1
IFNInterferons
IL-17Interleukin-17
IL-1βInterleukin-1β
IL-37Interleukin-37
IL-7Interleukin-7
KSHVKaposi’s sarcoma-associated herpesvirus
LKB1Liver kinase B1
LMP1Latent membrane protein 1
MAPKMitogen-activated protein kinase
MMP-8Matrix metalloproteinase-8
mTORMammalian target of rapamycin
mTORC1Mammalian target of rapamycin complex 1
NADPHNicotinamide adenine dinucleotide phosphate
NF-κBNuclear factor kappa-light-chain-enhancer of activated B cells
PPARGC1APeroxisome proliferator-activated receptor-gamma, coactivator 1α
ROSReactive oxygen species
RVFVRift Valley Fever Virus
SIRT1Sirtuin 1
SIVSimian immunodeficiency virus
SNARKSucrose-non-fermenting protein kinase 1/AMP-activated protein kinase-related protein kinase
STINGStimulator of IFN genes
TAK1Transforming growth factor (TGF)-β-activated kinase 1
TLRToll-like receptor
TSCTuberous sclerosis complex

References

  1. Hardie, D.G.; Ross, F.A.; Hawley, S.A. AMPK: A nutrient and energy sensor that maintains energy homeostasis. Nat. Rev. Mol. Cell Biol. 2012, 13, 251–262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Gowans, G.J.; Hardie, D.G. AMPK: A cellular energy sensor primarily regulated by AMP. Biochem. Soc. Trans. 2014, 42, 71–75. [Google Scholar] [CrossRef] [PubMed]
  3. Gowans, G.J.; Hawley, S.A.; Ross, F.A.; Hardie, D.G. AMP is a true physiological regulator of AMP-activated protein kinase by both allosteric activation and enhancing net phosphorylation. Cell Metab. 2013, 18, 556–566. [Google Scholar] [CrossRef] [PubMed]
  4. Shackelford, D.B.; Shaw, R.J. The LKB1-AMPK pathway: Metabolism and growth control in tumour suppression. Nat. Rev. Cancer 2009, 9, 563–575. [Google Scholar] [CrossRef] [PubMed]
  5. Faubert, B.; Boily, G.; Izreig, S.; Griss, T.; Samborska, B.; Dong, Z.; Dupuy, F.; Chambers, C.; Fuerth, B.J.; Viollet, B.; et al. AMPK is a negative regulator of the Warburg effect and suppresses tumor growth in vivo. Cell Metab. 2013, 17, 113–124. [Google Scholar] [CrossRef] [PubMed]
  6. Hardie, D.G. AMP-activated protein kinase: A cellular energy sensor with a key role in metabolic disorders and in cancer. Biochem. Soc. Trans. 2011, 39, 1–13. [Google Scholar] [CrossRef] [PubMed]
  7. Moreira, D.; Silvestre, R.; Cordeiro-da-Silva, A.; Estaquier, J.; Foretz, M.; Viollet, B. AMP-activated Protein Kinase as a Target for Pathogens: Friends or Foes? Curr. Drug Targets 2016, 17, 942–953. [Google Scholar] [CrossRef] [PubMed]
  8. Brunton, J.; Steele, S.; Ziehr, B.; Moorman, N.; Kawula, T. Feeding uninvited guests: MTOR and AMPK set the table for intracellular pathogens. PLoS Pathog. 2013, 9, e1003552. [Google Scholar] [CrossRef] [PubMed]
  9. Xie, N.; Yuan, K.; Zhou, L.; Wang, K.; Chen, H.N.; Lei, Y.; Lan, J.; Pu, Q.; Gao, W.; Zhang, L.; et al. PRKAA/AMPK restricts HBV replication through promotion of autophagic degradation. Autophagy 2016, 12, 1507–1520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Prantner, D.; Perkins, D.J.; Vogel, S.N. AMP-activated Kinase (AMPK) Promotes Innate Immunity and Antiviral Defense through Modulation of Stimulator of Interferon Genes (STING) Signaling. J. Biol. Chem. 2017, 292, 292–304. [Google Scholar] [CrossRef] [PubMed]
  11. Shin, D.M.; Yang, C.S.; Lee, J.Y.; Lee, S.J.; Choi, H.H.; Lee, H.M.; Yuk, J.M.; Harding, C.V.; Jo, E.K. Mycobacterium tuberculosis lipoprotein-induced association of TLR2 with protein kinase C zeta in lipid rafts contributes to reactive oxygen species-dependent inflammatory signalling in macrophages. Cell Microbiol. 2008, 10, 1893–1905. [Google Scholar] [CrossRef] [PubMed]
  12. Yang, C.S.; Kim, J.J.; Lee, H.M.; Jin, H.S.; Lee, S.H.; Park, J.H.; Kim, S.J.; Kim, J.M.; Han, Y.M.; Lee, M.S.; et al. The AMPK-PPARGC1A pathway is required for antimicrobial host defense through activation of autophagy. Autophagy 2014, 10, 785–802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Mankouri, J.; Harris, M. Viruses and the fuel sensor: The emerging link between AMPK and virus replication. Rev. Med. Virol. 2011, 21, 205–212. [Google Scholar] [CrossRef] [PubMed]
  14. Mesquita, I.; Moreira, D.; Sampaio-Marques, B.; Laforge, M.; Cordeiro-da-Silva, A.; Ludovico, P.; Estaquier, J.; Silvestre, R. AMPK in Pathogens. EXS 2016, 107, 287–323. [Google Scholar] [CrossRef] [PubMed]
  15. Kim, J.; Yang, G.; Kim, Y.; Kim, J.; Ha, J. AMPK activators: Mechanisms of action and physiological activities. Exp. Mol. Med. 2016, 48, e224. [Google Scholar] [CrossRef] [PubMed]
  16. Wu, J.; Puppala, D.; Feng, X.; Monetti, M.; Lapworth, A.L.; Geoghegan, K.F. Chemoproteomic analysis of intertissue and interspecies isoform diversity of AMP-activated protein kinase (AMPK). J. Biol. Chem. 2013, 288, 35904–35912. [Google Scholar] [CrossRef] [PubMed]
  17. Rajamohan, F.; Reyes, A.R.; Frisbie, R.K.; Hoth, L.R.; Sahasrabudhe, P.; Magyar, R.; Landro, J.A.; Withka, J.M.; Caspers, N.L.; Calabrese, M.F.; et al. Probing the enzyme kinetics, allosteric modulation and activation of alpha1- and alpha2-subunit-containing AMP-activated protein kinase (AMPK) heterotrimeric complexes by pharmacological and physiological activators. Biochem. J. 2016, 473, 581–592. [Google Scholar] [CrossRef] [PubMed]
  18. Woods, A.; Johnstone, S.R.; Dickerson, K.; Leiper, F.C.; Fryer, L.G.; Neumann, D.; Schlattner, U.; Wallimann, T.; Carlson, M.; Carling, D. LKB1 is the upstream kinase in the AMP-activated protein kinase cascade. Curr. Biol. 2003, 13, 2004–2008. [Google Scholar] [CrossRef] [PubMed]
  19. Carling, D.; Aguan, K.; Woods, A.; Verhoeven, A.J.; Beri, R.K.; Brennan, C.H.; Sidebottom, C.; Davison, M.D.; Scott, J. Mammalian AMP-activated protein kinase is homologous to yeast and plant protein kinases involved in the regulation of carbon metabolism. J. Biol. Chem. 1994, 269, 11442–11448. [Google Scholar] [PubMed]
  20. Hawley, S.A.; Pan, D.A.; Mustard, K.J.; Ross, L.; Bain, J.; Edelman, A.M.; Frenguelli, B.G.; Hardie, D.G. Calmodulin-dependent protein kinase kinase-beta is an alternative upstream kinase for AMP-activated protein kinase. Cell. Metab. 2005, 2, 9–19. [Google Scholar] [CrossRef] [PubMed]
  21. McBride, A.; Ghilagaber, S.; Nikolaev, A.; Hardie, D.G. The glycogen-binding domain on the AMPK beta subunit allows the kinase to act as a glycogen sensor. Cell. Metab. 2009, 9, 23–34. [Google Scholar] [CrossRef] [PubMed]
  22. Viana, R.; Towler, M.C.; Pan, D.A.; Carling, D.; Viollet, B.; Hardie, D.G.; Sanz, P. A conserved sequence immediately N-terminal to the Bateman domains in AMP-activated protein kinase gamma subunits is required for the interaction with the beta subunits. J. Biol. Chem. 2007, 282, 16117–16125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Xiao, B.; Heath, R.; Saiu, P.; Leiper, F.C.; Leone, P.; Jing, C.; Walker, P.A.; Haire, L.; Eccleston, J.F.; Davis, C.T.; et al. Structural basis for AMP binding to mammalian AMP-activated protein kinase. Nature 2007, 449, 496–500. [Google Scholar] [CrossRef] [PubMed]
  24. Steinberg, G.R.; Kemp, B.E. AMPK in Health and Disease. Physiol. Rev. 2009, 89, 1025–1078. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Viollet, B.; Horman, S.; Leclerc, J.; Lantier, L.; Foretz, M.; Billaud, M.; Giri, S.; Andreelli, F. AMPK inhibition in health and disease. Crit. Rev. Biochem. Mol. Biol. 2010, 45, 276–295. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Kim, T.S.; Shin, Y.H.; Lee, H.M.; Kim, J.K.; Choe, J.H.; Jang, J.C.; Um, S.; Jin, H.S.; Komatsu, M.; Cha, G.H.; et al. Ohmyungsamycins promote antimicrobial responses through autophagy activation via AMP-activated protein kinase pathway. Sci. Rep. 2017, 7, 3431. [Google Scholar] [CrossRef] [PubMed]
  27. Mankouri, J.; Tedbury, P.R.; Gretton, S.; Hughes, M.E.; Griffin, S.D.; Dallas, M.L.; Green, K.A.; Hardie, D.G.; Peers, C.; Harris, M. Enhanced hepatitis C virus genome replication and lipid accumulation mediated by inhibition of AMP-activated protein kinase. Proc. Natl. Acad. Sci. USA 2010, 107, 11549–11554. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Huang, H.; Kang, R.; Wang, J.; Luo, G.; Yang, W.; Zhao, Z. Hepatitis C virus inhibits AKT-tuberous sclerosis complex (TSC), the mechanistic target of rapamycin (MTOR) pathway, through endoplasmic reticulum stress to induce autophagy. Autophagy 2013, 9, 175–195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Yu, J.W.; Sun, L.J.; Liu, W.; Zhao, Y.H.; Kang, P.; Yan, B.Z. Hepatitis C virus core protein induces hepatic metabolism disorders through down-regulation of the SIRT1-AMPK signaling pathway. Int. J. Infect. Dis. 2013, 17, e539–e545. [Google Scholar] [CrossRef] [PubMed]
  30. Tsai, W.L.; Chang, T.H.; Sun, W.C.; Chan, H.H.; Wu, C.C.; Hsu, P.I.; Cheng, J.S.; Yu, M.L. Metformin activates type I interferon signaling against HCV via activation of adenosine monophosphate-activated protein kinase. Oncotarget 2017, 8, 91928–91937. [Google Scholar] [CrossRef] [PubMed]
  31. Nakashima, K.; Takeuchi, K.; Chihara, K.; Hotta, H.; Sada, K. Inhibition of hepatitis C virus replication through adenosine monophosphate-activated protein kinase-dependent and -independent pathways. Microbiol. Immunol. 2011, 55, 774–782. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Goto, K.; Lin, W.; Zhang, L.; Jilg, N.; Shao, R.X.; Schaefer, E.A.; Zhao, H.; Fusco, D.N.; Peng, L.F.; Kato, N.; et al. The AMPK-related kinase SNARK regulates hepatitis C virus replication and pathogenesis through enhancement of TGF-beta signaling. J. Hepatol. 2013, 59, 942–948. [Google Scholar] [CrossRef] [PubMed]
  33. Douglas, D.N.; Pu, C.H.; Lewis, J.T.; Bhat, R.; Anwar-Mohamed, A.; Logan, M.; Lund, G.; Addison, W.R.; Lehner, R.; Kneteman, N.M. Oxidative Stress Attenuates Lipid Synthesis and Increases Mitochondrial Fatty Acid Oxidation in Hepatoma Cells Infected with Hepatitis C Virus. J. Biol. Chem. 2016, 291, 1974–1990. [Google Scholar] [CrossRef] [PubMed]
  34. Uematsu, T.; Fujita, T.; Nakaoka, H.J.; Hara, T.; Kobayashi, N.; Murakami, Y.; Seiki, M.; Sakamoto, T. Mint3/Apba3 depletion ameliorates severe murine influenza pneumonia and macrophage cytokine production in response to the influenza virus. Sci. Rep. 2016, 6, 37815. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Yoon, J.Y.; Kwon, H.H.; Min, S.U.; Thiboutot, D.M.; Suh, D.H. Epigallocatechin-3-gallate improves acne in humans by modulating intracellular molecular targets and inhibiting P. acnes. J. Investig. Dermatol. 2013, 133, 429–440. [Google Scholar] [CrossRef] [PubMed]
  36. Han, S.; Xu, J.; Guo, X.; Huang, M. Curcumin ameliorates severe influenza pneumonia via attenuating lung injury and regulating macrophage cytokines production. Clin. Exp. Pharmacol. Physiol. 2018, 45, 84–93. [Google Scholar] [CrossRef] [PubMed]
  37. Zhang, H.S.; Wu, T.C.; Sang, W.W.; Ruan, Z. EGCG inhibits Tat-induced LTR transactivation: Role of Nrf2, AKT, AMPK signaling pathway. Life Sci. 2012, 90, 747–754. [Google Scholar] [CrossRef] [PubMed]
  38. Wang, Z.Q.; Yu, Y.; Zhang, X.H.; Floyd, E.Z.; Cefalu, W.T. Human adenovirus 36 decreases fatty acid oxidation and increases de novo lipogenesis in primary cultured human skeletal muscle cells by promoting Cidec/FSP27 expression. Int. J. Obes. 2010, 34, 1355–1364. [Google Scholar] [CrossRef] [PubMed]
  39. Moser, T.S.; Schieffer, D.; Cherry, S. AMP-activated kinase restricts Rift Valley fever virus infection by inhibiting fatty acid synthesis. PLoS Pathog. 2012, 8, e1002661. [Google Scholar] [CrossRef] [PubMed]
  40. Leyton, L.; Hott, M.; Acuna, F.; Caroca, J.; Nunez, M.; Martin, C.; Zambrano, A.; Concha, M.I.; Otth, C. Nutraceutical activators of AMPK/Sirt1 axis inhibit viral production and protect neurons from neurodegenerative events triggered during HSV-1 infection. Virus Res. 2015, 205, 63–72. [Google Scholar] [CrossRef] [PubMed]
  41. Xie, W.; Wang, L.; Dai, Q.; Yu, H.; He, X.; Xiong, J.; Sheng, H.; Zhang, D.; Xin, R.; Qi, Y.; et al. Activation of AMPK restricts coxsackievirus B3 replication by inhibiting lipid accumulation. J. Mol. Cell Cardiol. 2015, 85, 155–167. [Google Scholar] [CrossRef] [PubMed]
  42. Lo, A.K.; Lo, K.W.; Ko, C.W.; Young, L.S.; Dawson, C.W. Inhibition of the LKB1-AMPK pathway by the Epstein-Barr virus-encoded LMP1 promotes proliferation and transformation of human nasopharyngeal epithelial cells. J. Pathol. 2013, 230, 336–346. [Google Scholar] [CrossRef] [PubMed]
  43. Cheng, F.; He, M.; Jung, J.U.; Lu, C.; Gao, S.J. Suppression of Kaposi’s Sarcoma-Associated Herpesvirus Infection and Replication by 5′-AMP-Activated Protein Kinase. J. Virol. 2016, 90, 6515–6525. [Google Scholar] [CrossRef] [PubMed]
  44. Green, V.A.; Pelkmans, L. A Systems Survey of Progressive Host-Cell Reorganization during Rotavirus Infection. Cell Host Microbe 2016, 20, 107–120. [Google Scholar] [CrossRef] [PubMed]
  45. Soto-Acosta, R.; Bautista-Carbajal, P.; Cervantes-Salazar, M.; Angel-Ambrocio, A.H.; Del Angel, R.M. DENV up-regulates the HMG-CoA reductase activity through the impairment of AMPK phosphorylation: A potential antiviral target. PLoS Pathog. 2017, 13, e1006257. [Google Scholar] [CrossRef] [PubMed]
  46. Jordan, T.X.; Randall, G. Dengue Virus Activates the AMP Kinase-mTOR Axis to Stimulate a Proviral Lipophagy. J. Virol. 2017, 91. [Google Scholar] [CrossRef] [PubMed]
  47. Bagga, S.; Rawat, S.; Ajenjo, M.; Bouchard, M.J. Hepatitis B virus (HBV) X protein-mediated regulation of hepatocyte metabolic pathways affects viral replication. Virology 2016, 498, 9–22. [Google Scholar] [CrossRef] [PubMed]
  48. Kondratowicz, A.S.; Hunt, C.L.; Davey, R.A.; Cherry, S.; Maury, W.J. AMP-activated protein kinase is required for the macropinocytic internalization of ebolavirus. J. Virol. 2013, 87, 746–755. [Google Scholar] [CrossRef] [PubMed]
  49. Ji, W.T.; Lee, L.H.; Lin, F.L.; Wang, L.; Liu, H.J. AMP-activated protein kinase facilitates avian reovirus to induce mitogen-activated protein kinase (MAPK) p38 and MAPK kinase 3/6 signalling that is beneficial for virus replication. J. Gen. Virol. 2009, 90, 3002–3009. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Chi, P.I.; Huang, W.R.; Lai, I.H.; Cheng, C.Y.; Liu, H.J. The p17 nonstructural protein of avian reovirus triggers autophagy enhancing virus replication via activation of phosphatase and tensin deleted on chromosome 10 (PTEN) and AMP-activated protein kinase (AMPK), as well as dsRNA-dependent protein kinase (PKR)/eIF2alpha signaling pathways. J. Biol. Chem. 2013, 288, 3571–3584. [Google Scholar] [CrossRef] [PubMed]
  51. Martin, C.; Leyton, L.; Arancibia, Y.; Cuevas, A.; Zambrano, A.; Concha, M.I.; Otth, C. Modulation of the AMPK/Sirt1 axis during neuronal infection by herpes simplex virus type 1. J. Alzheimers Dis. 2014, 42, 301–312. [Google Scholar] [CrossRef] [PubMed]
  52. Samikkannu, T.; Atluri, V.S.; Nair, M.P. HIV and Cocaine Impact Glial Metabolism: Energy Sensor AMP-activated protein kinase Role in Mitochondrial Biogenesis and Epigenetic Remodeling. Sci. Rep. 2016, 6, 31784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Terry, L.J.; Vastag, L.; Rabinowitz, J.D.; Shenk, T. Human kinome profiling identifies a requirement for AMP-activated protein kinase during human cytomegalovirus infection. Proc. Natl. Acad. Sci. USA 2012, 109, 3071–3076. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Hutterer, C.; Wandinger, S.K.; Wagner, S.; Muller, R.; Stamminger, T.; Zeittrager, I.; Godl, K.; Baumgartner, R.; Strobl, S.; Marschall, M. Profiling of the kinome of cytomegalovirus-infected cells reveals the functional importance of host kinases Aurora A, ABL and AMPK. Antivir. Res. 2013, 99, 139–148. [Google Scholar] [CrossRef] [PubMed]
  55. McArdle, J.; Moorman, N.J.; Munger, J. HCMV targets the metabolic stress response through activation of AMPK whose activity is important for viral replication. PLoS Pathog. 2012, 8, e1002502. [Google Scholar] [CrossRef] [PubMed]
  56. Mukhopadhyay, R.; Venkatadri, R.; Katsnelson, J.; Arav-Boger, R. Digitoxin Suppresses Human Cytomegalovirus Replication via Na+, K+/ATPase alpha1 Subunit-Dependent AMP-Activated Protein Kinase and Autophagy Activation. J. Virol. 2018, 92. [Google Scholar] [CrossRef] [PubMed]
  57. Seo, J.Y.; Yaneva, R.; Hinson, E.R.; Cresswell, P. Human cytomegalovirus directly induces the antiviral protein viperin to enhance infectivity. Science 2011, 332, 1093–1097. [Google Scholar] [CrossRef] [PubMed]
  58. Seo, J.Y.; Cresswell, P. Viperin regulates cellular lipid metabolism during human cytomegalovirus infection. PLoS Pathog. 2013, 9, e1003497. [Google Scholar] [CrossRef] [PubMed]
  59. Li, M.; Li, J.; Zeng, R.; Yang, J.; Liu, J.; Zhang, Z.; Song, X.; Yao, Z.; Ma, C.; Li, W.; et al. Respiratory Syncytial Virus Replication Is Promoted by Autophagy-Mediated Inhibition of Apoptosis. J. Virol. 2018, 92. [Google Scholar] [CrossRef] [PubMed]
  60. Lv, S.; Xu, Q.Y.; Sun, E.C.; Zhang, J.K.; Wu, D.L. Dissection and integration of the autophagy signaling network initiated by bluetongue virus infection: Crucial candidates ERK1/2, Akt and AMPK. Sci. Rep. 2016, 6, 23130. [Google Scholar] [CrossRef] [PubMed]
  61. Liu, J.; Wang, H.; Gu, J.; Deng, T.; Yuan, Z.; Hu, B.; Xu, Y.; Yan, Y.; Zan, J.; Liao, M.; et al. BECN1-dependent CASP2 incomplete autophagy induction by binding to rabies virus phosphoprotein. Autophagy 2017, 13, 739–753. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Subramanian, G.; Kuzmanovic, T.; Zhang, Y.; Peter, C.B.; Veleeparambil, M.; Chakravarti, R.; Sen, G.C.; Chattopadhyay, S. A new mechanism of interferon’s antiviral action: Induction of autophagy, essential for paramyxovirus replication, is inhibited by the interferon stimulated gene, TDRD7. PLoS Pathog. 2018, 14, e1006877. [Google Scholar] [CrossRef] [PubMed]
  63. McFadden, K.; Hafez, A.Y.; Kishton, R.; Messinger, J.E.; Nikitin, P.A.; Rathmell, J.C.; Luftig, M.A. Metabolic stress is a barrier to Epstein-Barr virus-mediated B-cell immortalization. Proc. Natl. Acad. Sci. USA 2016, 113, E782–E790. [Google Scholar] [CrossRef] [PubMed]
  64. Burke, J.D.; Platanias, L.C.; Fish, E.N. Beta interferon regulation of glucose metabolism is PI3K/Akt dependent and important for antiviral activity against coxsackievirus B3. J. Virol. 2014, 88, 3485–3495. [Google Scholar] [CrossRef] [PubMed]
  65. Zhang, C.; Feng, S.; Zhang, W.; Chen, N.; Hegazy, A.M.; Chen, W.; Liu, X.; Zhao, L.; Li, J.; Lin, L.; et al. MicroRNA miR-214 Inhibits Snakehead Vesiculovirus Replication by Promoting IFN-alpha Expression via Targeting Host Adenosine 5′-Monophosphate-Activated Protein Kinase. Front. Immunol. 2017, 8, 1775. [Google Scholar] [CrossRef] [PubMed]
  66. Francione, L.; Smith, P.K.; Accari, S.L.; Taylor, P.E.; Bokko, P.B.; Bozzaro, S.; Beech, P.L.; Fisher, P.R. Legionella pneumophila multiplication is enhanced by chronic AMPK signalling in mitochondrially diseased Dictyostelium cells. Dis. Model Mech. 2009, 2, 479–489. [Google Scholar] [CrossRef] [PubMed]
  67. Kajiwara, C.; Kusaka, Y.; Kimura, S.; Yamaguchi, T.; Nanjo, Y.; Ishii, Y.; Udono, H.; Standiford, T.J.; Tateda, K. Metformin Mediates Protection against Legionella Pneumonia through Activation of AMPK and Mitochondrial Reactive Oxygen Species. J. Immunol. 2018, 200, 623–631. [Google Scholar] [CrossRef] [PubMed]
  68. Ganesan, R.; Hos, N.J.; Gutierrez, S.; Fischer, J.; Stepek, J.M.; Daglidu, E.; Kronke, M.; Robinson, N. Salmonella Typhimurium disrupts Sirt1/AMPK checkpoint control of mTOR to impair autophagy. PLoS Pathog. 2017, 13, e1006227. [Google Scholar] [CrossRef] [PubMed]
  69. Liu, W.; Jiang, Y.; Sun, J.; Geng, S.; Pan, Z.; Prinz, R.A.; Wang, C.; Sun, J.; Jiao, X.; Xu, X. Activation of TGF-beta-activated kinase 1 (TAK1) restricts Salmonella Typhimurium growth by inducing AMPK activation and autophagy. Cell Death Dis. 2018, 9, 570. [Google Scholar] [CrossRef] [PubMed]
  70. Tattoli, I.; Sorbara, M.T.; Philpott, D.J.; Girardin, S.E. Bacterial autophagy: The trigger, the target and the timing. Autophagy 2012, 8, 1848–1850. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Neumann, D. Is TAK1 a Direct Upstream Kinase of AMPK? Int. J. Mol. Sci. 2018, 19, 2412. [Google Scholar] [CrossRef] [PubMed]
  72. Liu, N.; Li, Y.; Dong, C.; Xu, X.; Wei, P.; Sun, W.; Peng, Q. Inositol-Requiring Enzyme 1-Dependent Activation of AMPK Promotes Brucella abortus Intracellular Growth. J. Bacteriol. 2016, 198, 986–993. [Google Scholar] [CrossRef] [PubMed]
  73. Liang, Y.D.; Bai, W.J.; Li, C.G.; Xu, L.H.; Wei, H.X.; Pan, H.; He, X.H.; Ouyang, D.Y. Piperine Suppresses Pyroptosis and Interleukin-1beta Release upon ATP Triggering and Bacterial Infection. Front. Pharmacol. 2016, 7, 390. [Google Scholar] [CrossRef] [PubMed]
  74. Yang, W.; Dierking, K.; Esser, D.; Tholey, A.; Leippe, M.; Rosenstiel, P.; Schulenburg, H. Overlapping and unique signatures in the proteomic and transcriptomic responses of the nematode Caenorhabditis elegans toward pathogenic Bacillus thuringiensis. Dev. Comp. Immunol. 2015, 51, 1–9. [Google Scholar] [CrossRef] [PubMed]
  75. Lv, G.; Zhu, H.; Zhou, F.; Lin, Z.; Lin, G.; Li, C. AMP-activated protein kinase activation protects gastric epithelial cells from Helicobacter pylori-induced apoptosis. Biochem. Biophys. Res. Commun. 2014, 453, 13–18. [Google Scholar] [CrossRef] [PubMed]
  76. Zhao, H.; Zhu, H.; Lin, Z.; Lin, G.; Lv, G. Compound 13, an alpha1-selective small molecule activator of AMPK, inhibits Helicobacter pylori-induced oxidative stresses and gastric epithelial cell apoptosis. Biochem. Biophys. Res. Commun. 2015, 463, 510–517. [Google Scholar] [CrossRef] [PubMed]
  77. Patkee, W.R.; Carr, G.; Baker, E.H.; Baines, D.L.; Garnett, J.P. Metformin prevents the effects of Pseudomonas aeruginosa on airway epithelial tight junctions and restricts hyperglycaemia-induced bacterial growth. J. Cell Mol. Med. 2016, 20, 758–764. [Google Scholar] [CrossRef] [PubMed]
  78. Kumar, A.; Giri, S.; Kumar, A. 5-Aminoimidazole-4-carboxamide ribonucleoside-mediated adenosine monophosphate-activated protein kinase activation induces protective innate responses in bacterial endophthalmitis. Cell Microbiol. 2016, 18, 1815–1830. [Google Scholar] [CrossRef] [PubMed]
  79. Wang, Y.; Zhang, K.; Shi, X.; Wang, C.; Wang, F.; Fan, J.; Shen, F.; Xu, J.; Bao, W.; Liu, M.; et al. Critical role of bacterial isochorismatase in the autophagic process induced by Acinetobacter baumannii in mammalian cells. FASEB J. 2016, 30, 3563–3577. [Google Scholar] [CrossRef] [PubMed]
  80. Singhal, A.; Jie, L.; Kumar, P.; Hong, G.S.; Leow, M.K.; Paleja, B.; Tsenova, L.; Kurepina, N.; Chen, J.; Zolezzi, F.; et al. Metformin as adjunct antituberculosis therapy. Sci. Transl. Med. 2014, 6, 263ra159. [Google Scholar] [CrossRef] [PubMed]
  81. Yuk, J.M.; Shin, D.M.; Lee, H.M.; Yang, C.S.; Jin, H.S.; Kim, K.K.; Lee, Z.W.; Lee, S.H.; Kim, J.M.; Jo, E.K. Vitamin D3 induces autophagy in human monocytes/macrophages via cathelicidin. Cell Host Microbe 2009, 6, 231–243. [Google Scholar] [CrossRef] [PubMed]
  82. Rekha, R.S.; Rao Muvva, S.S.; Wan, M.; Raqib, R.; Bergman, P.; Brighenti, S.; Gudmundsson, G.H.; Agerberth, B. Phenylbutyrate induces LL-37-dependent autophagy and intracellular killing of Mycobacterium tuberculosis in human macrophages. Autophagy 2015, 11, 1688–1699. [Google Scholar] [CrossRef] [PubMed]
  83. Kim, J.K.; Kim, Y.S.; Lee, H.M.; Jin, H.S.; Neupane, C.; Kim, S.; Lee, S.H.; Min, J.J.; Sasai, M.; Jeong, J.H.; et al. GABAergic signaling linked to autophagy enhances host protection against intracellular bacterial infections. Nat. Commun. 2018, 9, 4184. [Google Scholar] [CrossRef] [PubMed]
  84. Ong, C.W.; Elkington, P.T.; Brilha, S.; Ugarte-Gil, C.; Tome-Esteban, M.T.; Tezera, L.B.; Pabisiak, P.J.; Moores, R.C.; Sathyamoorthy, T.; Patel, V.; et al. Neutrophil-Derived MMP-8 Drives AMPK-Dependent Matrix Destruction in Human Pulmonary Tuberculosis. PLoS Pathog. 2015, 11, e1004917. [Google Scholar] [CrossRef] [PubMed]
  85. Restrepo, B.I. Metformin: Candidate host-directed therapy for tuberculosis in diabetes and non-diabetes patients. Tuberculosis (Edinb) 2016, 101S, S69–S72. [Google Scholar] [CrossRef] [PubMed]
  86. Ouimet, M.; Koster, S.; Sakowski, E.; Ramkhelawon, B.; van Solingen, C.; Oldebeken, S.; Karunakaran, D.; Portal-Celhay, C.; Sheedy, F.J.; Ray, T.D.; et al. Mycobacterium tuberculosis induces the miR-33 locus to reprogram autophagy and host lipid metabolism. Nat. Immunol. 2016, 17, 677–686. [Google Scholar] [CrossRef] [PubMed]
  87. Kim, J.K.; Lee, H.M.; Park, K.S.; Shin, D.M.; Kim, T.S.; Kim, Y.S.; Suh, H.W.; Kim, S.Y.; Kim, I.S.; Kim, J.M.; et al. MIR144* inhibits antimicrobial responses against Mycobacterium tuberculosis in human monocytes and macrophages by targeting the autophagy protein DRAM2. Autophagy 2017, 13, 423–441. [Google Scholar] [CrossRef] [PubMed]
  88. Dorhoi, A.; Kaufmann, S.H. Pathology and immune reactivity: Understanding multidimensionality in pulmonary tuberculosis. Semin. Immunopathol. 2016, 38, 153–166. [Google Scholar] [CrossRef] [PubMed]
  89. Nieves, W.; Hung, L.Y.; Oniskey, T.K.; Boon, L.; Foretz, M.; Viollet, B.; Herbert, D.R. Myeloid-Restricted AMPKalpha1 Promotes Host Immunity and Protects against IL-12/23p40-Dependent Lung Injury during Hookworm Infection. J. Immunol. 2016, 196, 4632–4640. [Google Scholar] [CrossRef] [PubMed]
  90. Moreira, D.; Rodrigues, V.; Abengozar, M.; Rivas, L.; Rial, E.; Laforge, M.; Li, X.; Foretz, M.; Viollet, B.; Estaquier, J.; et al. Leishmania infantum modulates host macrophage mitochondrial metabolism by hijacking the SIRT1-AMPK axis. PLoS Pathog. 2015, 11, e1004684. [Google Scholar] [CrossRef] [PubMed]
  91. Zhu, J.; Zhang, W.; Zhang, L.; Xu, L.; Chen, X.; Zhou, S.; Xu, Z.; Xiao, M.; Bai, H.; Liu, F.; et al. IL-7 suppresses macrophage autophagy and promotes liver pathology in Schistosoma japonicum-infected mice. J. Cell. Mol. Med. 2018, 22, 3353–3363. [Google Scholar] [CrossRef] [PubMed]
  92. Ruivo, M.T.G.; Vera, I.M.; Sales-Dias, J.; Meireles, P.; Gural, N.; Bhatia, S.N.; Mota, M.M.; Mancio-Silva, L. Host AMPK Is a Modulator of Plasmodium Liver Infection. Cell. Rep. 2016, 16, 2539–2545. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Vilar-Pereira, G.; Carneiro, V.C.; Mata-Santos, H.; Vicentino, A.R.; Ramos, I.P.; Giarola, N.L.; Feijo, D.F.; Meyer-Fernandes, J.R.; Paula-Neto, H.A.; Medei, E.; et al. Resveratrol Reverses Functional Chagas Heart Disease in Mice. PLoS Pathog. 2016, 12, e1005947. [Google Scholar] [CrossRef] [PubMed]
  94. Pandey, A.; Ding, S.L.; Qin, Q.M.; Gupta, R.; Gomez, G.; Lin, F.; Feng, X.; Fachini da Costa, L.; Chaki, S.P.; Katepalli, M.; et al. Global Reprogramming of Host Kinase Signaling in Response to Fungal Infection. Cell Host Microbe 2017, 21, 637–649. [Google Scholar] [CrossRef] [PubMed]
  95. Andris, F.; Leo, O. AMPK in lymphocyte metabolism and function. Int. Rev. Immunol. 2015, 34, 67–81. [Google Scholar] [CrossRef] [PubMed]
  96. Vincent, E.E.; Coelho, P.P.; Blagih, J.; Griss, T.; Viollet, B.; Jones, R.G. Differential effects of AMPK agonists on cell growth and metabolism. Oncogene 2015, 34, 3627–3639. [Google Scholar] [CrossRef] [PubMed]
  97. Bain, J.; Plater, L.; Elliott, M.; Shpiro, N.; Hastie, C.J.; McLauchlan, H.; Klevernic, I.; Arthur, J.S.; Alessi, D.R.; Cohen, P. The selectivity of protein kinase inhibitors: A further update. Biochem. J. 2007, 408, 297–315. [Google Scholar] [CrossRef] [PubMed]
  98. Myers, R.W.; Guan, H.P.; Ehrhart, J.; Petrov, A.; Prahalada, S.; Tozzo, E.; Yang, X.; Kurtz, M.M.; Trujillo, M.; Gonzalez Trotter, D.; et al. Systemic pan-AMPK activator MK-8722 improves glucose homeostasis but induces cardiac hypertrophy. Science 2017, 357, 507–511. [Google Scholar] [CrossRef] [PubMed]
  99. Ngoei, K.R.W.; Langendorf, C.G.; Ling, N.X.Y.; Hoque, A.; Varghese, S.; Camerino, M.A.; Walker, S.R.; Bozikis, Y.E.; Dite, T.A.; Ovens, A.J.; et al. Structural Determinants for Small-Molecule Activation of Skeletal Muscle AMPK alpha2beta2gamma1 by the Glucose Importagog SC4. Cell Chem. Biol. 2018, 25, 728–737. [Google Scholar] [CrossRef] [PubMed]
  100. Dite, T.A.; Langendorf, C.G.; Hoque, A.; Galic, S.; Rebello, R.J.; Ovens, A.J.; Lindqvist, L.M.; Ngoei, K.R.W.; Ling, N.X.Y.; Furic, L.; et al. AMP-activated protein kinase selectively inhibited by the type II inhibitor SBI-0206965. J. Biol. Chem. 2018, 293, 8874–8885. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Shintani, Y.; Kapoor, A.; Kaneko, M.; Smolenski, R.T.; D’Acquisto, F.; Coppen, S.R.; Harada-Shoji, N.; Lee, H.J.; Thiemermann, C.; Takashima, S.; et al. TLR9 mediates cellular protection by modulating energy metabolism in cardiomyocytes and neurons. Proc. Natl. Acad. Sci. USA 2013, 110, 5109–5114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Hoogendijk, A.J.; Pinhancos, S.S.; van der Poll, T.; Wieland, C.W. AMP-activated protein kinase activation by 5-aminoimidazole-4-carbox-amide-1-beta-D-ribofuranoside (AICAR) reduces lipoteichoic acid-induced lung inflammation. J. Biol. Chem. 2013, 288, 7047–7052. [Google Scholar] [CrossRef] [PubMed]
  103. O’Sullivan, T.E.; Johnson, L.R.; Kang, H.H.; Sun, J.C. BNIP3- and BNIP3L-Mediated Mitophagy Promotes the Generation of Natural Killer Cell Memory. Immunity 2015, 43, 331–342. [Google Scholar] [CrossRef] [PubMed]
  104. Muller-Durovic, B.; Lanna, A.; Covre, L.P.; Mills, R.S.; Henson, S.M.; Akbar, A.N. Killer Cell Lectin-like Receptor G1 Inhibits NK Cell Function through Activation of Adenosine 5′-Monophosphate-Activated Protein Kinase. J. Immunol. 2016, 197, 2891–2899. [Google Scholar] [CrossRef] [PubMed]
  105. Webb, T.J.; Carey, G.B.; East, J.E.; Sun, W.; Bollino, D.R.; Kimball, A.S.; Brutkiewicz, R.R. Alterations in cellular metabolism modulate CD1d-mediated NKT-cell responses. Pathog. Dis. 2016, 74. [Google Scholar] [CrossRef] [PubMed]
  106. Park, D.W.; Jiang, S.; Tadie, J.M.; Stigler, W.S.; Gao, Y.; Deshane, J.; Abraham, E.; Zmijewski, J.W. Activation of AMPK enhances neutrophil chemotaxis and bacterial killing. Mol. Med. 2013, 19, 387–398. [Google Scholar] [CrossRef] [PubMed]
  107. Zhao, X.; Zmijewski, J.W.; Lorne, E.; Liu, G.; Park, Y.J.; Tsuruta, Y.; Abraham, E. Activation of AMPK attenuates neutrophil proinflammatory activity and decreases the severity of acute lung injury. Am. J. Physiol. Lung Cell Mol. Physiol. 2008, 295, L497–L504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Bergheim, I.; Luyendyk, J.P.; Steele, C.; Russell, G.K.; Guo, L.; Roth, R.A.; Arteel, G.E. Metformin prevents endotoxin-induced liver injury after partial hepatectomy. J. Pharmacol. Exp. Ther. 2006, 316, 1053–1061. [Google Scholar] [CrossRef] [PubMed]
  109. Escobar, D.A.; Botero-Quintero, A.M.; Kautza, B.C.; Luciano, J.; Loughran, P.; Darwiche, S.; Rosengart, M.R.; Zuckerbraun, B.S.; Gomez, H. Adenosine monophosphate-activated protein kinase activation protects against sepsis-induced organ injury and inflammation. J. Surg. Res. 2015, 194, 262–272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Hattori, Y.; Suzuki, K.; Hattori, S.; Kasai, K. Metformin inhibits cytokine-induced nuclear factor κB activation via AMP-activated protein kinase activation in vascular endothelial cells. Hypertension 2006, 47, 1183–1188. [Google Scholar] [CrossRef] [PubMed]
  111. Zmijewski, J.W.; Lorne, E.; Zhao, X.; Tsuruta, Y.; Sha, Y.; Liu, G.; Siegal, G.P.; Abraham, E. Mitochondrial respiratory complex I regulates neutrophil activation and severity of lung injury. Am. J. Respir. Crit. Care Med. 2008, 178, 168–179. [Google Scholar] [CrossRef] [PubMed]
  112. Sag, D.; Carling, D.; Stout, R.D.; Suttles, J. Adenosine 5′-monophosphate-activated protein kinase promotes macrophage polarization to an anti-inflammatory functional phenotype. J. Immunol. 2008, 181, 8633–8641. [Google Scholar] [CrossRef] [PubMed]
  113. Xing, J.; Wang, Q.; Coughlan, K.; Viollet, B.; Moriasi, C.; Zou, M.H. Inhibition of AMP-activated protein kinase accentuates lipopolysaccharide-induced lung endothelial barrier dysfunction and lung injury in vivo. Am. J. Pathol. 2013, 182, 1021–1030. [Google Scholar] [CrossRef] [PubMed]
  114. Jeong, H.W.; Hsu, K.C.; Lee, J.W.; Ham, M.; Huh, J.Y.; Shin, H.J.; Kim, W.S.; Kim, J.B. Berberine suppresses proinflammatory responses through AMPK activation in macrophages. Am. J. Physiol. Endocrinol. Metab. 2009, 296, E955–E964. [Google Scholar] [CrossRef] [PubMed]
  115. Vaez, H.; Rameshrad, M.; Najafi, M.; Barar, J.; Barzegari, A.; Garjani, A. Cardioprotective effect of metformin in lipopolysaccharide-induced sepsis via suppression of toll-like receptor 4 (TLR4) in heart. Eur. J. Pharmacol. 2016, 772, 115–123. [Google Scholar] [CrossRef] [PubMed]
  116. Kim, J.; Kwak, H.J.; Cha, J.Y.; Jeong, Y.S.; Rhee, S.D.; Kim, K.R.; Cheon, H.G. Metformin suppresses lipopolysaccharide (LPS)-induced inflammatory response in murine macrophages via activating transcription factor-3 (ATF-3) induction. J. Biol. Chem. 2014, 289, 23246–23255. [Google Scholar] [CrossRef] [PubMed]
  117. Mulchandani, N.; Yang, W.L.; Khan, M.M.; Zhang, F.; Marambaud, P.; Nicastro, J.; Coppa, G.F.; Wang, P. Stimulation of Brain AMP-Activated Protein Kinase Attenuates Inflammation and Acute Lung Injury in Sepsis. Mol. Med. 2015, 21, 637–644. [Google Scholar] [CrossRef] [PubMed]
  118. Liu, Z.; Bone, N.; Jiang, S.; Park, D.W.; Tadie, J.M.; Deshane, J.; Rodriguez, C.A.; Pittet, J.F.; Abraham, E.; Zmijewski, J.W. AMP-Activated Protein Kinase and Glycogen Synthase Kinase 3beta Modulate the Severity of Sepsis-Induced Lung Injury. Mol. Med. 2016, 21, 937–950. [Google Scholar] [CrossRef] [PubMed]
  119. Chen, J.; Lai, J.; Yang, L.; Ruan, G.; Chaugai, S.; Ning, Q.; Chen, C.; Wang, D.W. Trimetazidine prevents macrophage-mediated septic myocardial dysfunction via activation of the histone deacetylase sirtuin 1. Br. J. Pharmacol. 2016, 173, 545–561. [Google Scholar] [CrossRef] [PubMed]
  120. Liu, X.; Wang, N.; Fan, S.; Zheng, X.; Yang, Y.; Zhu, Y.; Lu, Y.; Chen, Q.; Zhou, H.; Zheng, J. The citrus flavonoid naringenin confers protection in a murine endotoxaemia model through AMPK-ATF3-dependent negative regulation of the TLR4 signalling pathway. Sci. Rep. 2016, 6, 39735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Zhan, W.; Kang, Y.; Chen, N.; Mao, C.; Kang, Y.; Shang, J. Halofuginone ameliorates inflammation in severe acute hepatitis B virus (HBV)-infected SD rats through AMPK activation. Drug Des. Dev. Ther. 2017, 11, 2947–2955. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Bae, H.B.; Zmijewski, J.W.; Deshane, J.S.; Tadie, J.M.; Chaplin, D.D.; Takashima, S.; Abraham, E. AMP-activated protein kinase enhances the phagocytic ability of macrophages and neutrophils. FASEB J. 2011, 25, 4358–4368. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Li, L.; Xu, S.; Guo, T.; Gong, S.; Zhang, C. Effect of dapagliflozin on intestinal flora in MafA-deficient mice. Curr. Pharm. Des. 2018. [Google Scholar] [CrossRef] [PubMed]
  124. Rousset, C.I.; Leiper, F.C.; Kichev, A.; Gressens, P.; Carling, D.; Hagberg, H.; Thornton, C. A dual role for AMP-activated protein kinase (AMPK) during neonatal hypoxic-ischaemic brain injury in mice. J. Neurochem. 2015, 133, 242–252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Li, C.G.; Yan, L.; Jing, Y.Y.; Xu, L.H.; Liang, Y.D.; Wei, H.X.; Hu, B.; Pan, H.; Zha, Q.B.; Ouyang, D.Y.; et al. Berberine augments ATP-induced inflammasome activation in macrophages by enhancing AMPK signaling. Oncotarget 2017, 8, 95–109. [Google Scholar] [CrossRef] [PubMed]
  126. Zha, Q.B.; Wei, H.X.; Li, C.G.; Liang, Y.D.; Xu, L.H.; Bai, W.J.; Pan, H.; He, X.H.; Ouyang, D.Y. ATP-Induced Inflammasome Activation and Pyroptosis Is Regulated by AMP-Activated Protein Kinase in Macrophages. Front. Immunol. 2016, 7, 597. [Google Scholar] [CrossRef] [PubMed]
  127. Wen, H.; Gris, D.; Lei, Y.; Jha, S.; Zhang, L.; Huang, M.T.; Brickey, W.J.; Ting, J.P. Fatty acid-induced NLRP3-ASC inflammasome activation interferes with insulin signaling. Nat. Immunol. 2011, 12, 408–415. [Google Scholar] [CrossRef] [PubMed]
  128. Lee, H.M.; Kim, J.J.; Kim, H.J.; Shong, M.; Ku, B.J.; Jo, E.K. Upregulated NLRP3 inflammasome activation in patients with type 2 diabetes. Diabetes 2013, 62, 194–204. [Google Scholar] [CrossRef] [PubMed]
  129. Bullon, P.; Alcocer-Gomez, E.; Carrion, A.M.; Marin-Aguilar, F.; Garrido-Maraver, J.; Roman-Malo, L.; Ruiz-Cabello, J.; Culic, O.; Ryffel, B.; Apetoh, L.; et al. AMPK Phosphorylation Modulates Pain by Activation of NLRP3 Inflammasome. Antioxid. Redox. Signal. 2016, 24, 157–170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Wang, Y.; Viollet, B.; Terkeltaub, R.; Liu-Bryan, R. AMP-activated protein kinase suppresses urate crystal-induced inflammation and transduces colchicine effects in macrophages. Ann. Rheum. Dis. 2016, 75, 286–294. [Google Scholar] [CrossRef] [PubMed]
  131. Blagih, J.; Coulombe, F.; Vincent, E.E.; Dupuy, F.; Galicia-Vazquez, G.; Yurchenko, E.; Raissi, T.C.; van der Windt, G.J.; Viollet, B.; Pearce, E.L.; et al. The energy sensor AMPK regulates T cell metabolic adaptation and effector responses in vivo. Immunity 2015, 42, 41–54. [Google Scholar] [CrossRef] [PubMed]
  132. Iseda, S.; Takahashi, N.; Poplimont, H.; Nomura, T.; Seki, S.; Nakane, T.; Nakamura, M.; Shi, S.; Ishii, H.; Furukawa, S.; et al. Biphasic CD8+ T-Cell Defense in Simian Immunodeficiency Virus Control by Acute-Phase Passive Neutralizing Antibody Immunization. J. Virol. 2016, 90, 6276–6290. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Domain structures of the 5′ AMP-activated protein kinase (AMPK) subunits and the mechanisms that regulate activation of AMPK signaling pathways. (A) Conserved domain structure of AMPK subunits consisting of a catalytic α subunit, scaffolding β subunit, and regulatory γ subunit. AID, autoinhibitory domain; CBM, carbohydrate-binding module; CBS, cystathionine-beta-synthase; CTD, C-terminal domain. (B) AMPK is activated by the upstream kinases LKB1, CAMKK2 and TAK1 associated with the canonical pathway (triggered by an increased cellular AMP/ATP ratio) or the non-canonical pathway (triggered by an increased intracellular Ca2+ concentration or infection/TLR activation). Activated AMPK modulates cellular homeostasis, such as energy metabolism and autophagy, and mitochondrial homeostasis. (black arrow indicate activation/increase; bar-headed red arrow indicates inhibition/decrease). CAMKK2, calcium/calmodulin-dependent kinase kinase 2; LKB1, liver kinase B1; TAK1, Transforming growth factor-β-activated kinase 1; TLR, Toll-like receptor.
Figure 1. Domain structures of the 5′ AMP-activated protein kinase (AMPK) subunits and the mechanisms that regulate activation of AMPK signaling pathways. (A) Conserved domain structure of AMPK subunits consisting of a catalytic α subunit, scaffolding β subunit, and regulatory γ subunit. AID, autoinhibitory domain; CBM, carbohydrate-binding module; CBS, cystathionine-beta-synthase; CTD, C-terminal domain. (B) AMPK is activated by the upstream kinases LKB1, CAMKK2 and TAK1 associated with the canonical pathway (triggered by an increased cellular AMP/ATP ratio) or the non-canonical pathway (triggered by an increased intracellular Ca2+ concentration or infection/TLR activation). Activated AMPK modulates cellular homeostasis, such as energy metabolism and autophagy, and mitochondrial homeostasis. (black arrow indicate activation/increase; bar-headed red arrow indicates inhibition/decrease). CAMKK2, calcium/calmodulin-dependent kinase kinase 2; LKB1, liver kinase B1; TAK1, Transforming growth factor-β-activated kinase 1; TLR, Toll-like receptor.
Ijms 19 03495 g001aIjms 19 03495 g001b
Figure 2. Multifaceted roles of AMPK in viral and bacterial infections. A variety of viruses and bacteria modulate host AMPK activity to promote their growth in host cells. Activation of the AMPK signaling pathway has been implicated in both beneficial antiviral (left upper) and detrimental proviral (right upper) responses. In addition, AMPK activation promotes the host response to infections by various bacteria (left lower) but, in some cases, promotes a detrimental response (right lower). The detailed mechanisms by which AMPK activation/inhibition affects infection outcomes are listed in Table 1, Table 2, Table 3 and Table 4.
Figure 2. Multifaceted roles of AMPK in viral and bacterial infections. A variety of viruses and bacteria modulate host AMPK activity to promote their growth in host cells. Activation of the AMPK signaling pathway has been implicated in both beneficial antiviral (left upper) and detrimental proviral (right upper) responses. In addition, AMPK activation promotes the host response to infections by various bacteria (left lower) but, in some cases, promotes a detrimental response (right lower). The detailed mechanisms by which AMPK activation/inhibition affects infection outcomes are listed in Table 1, Table 2, Table 3 and Table 4.
Ijms 19 03495 g002
Figure 3. Regulatory effects and underlying mechanisms of small-molecule AMPK activators on the innate immune and inflammatory responses. (red upward arrows indicate activation/increase and blue downward arrows indicate inhibition/decrease)
Figure 3. Regulatory effects and underlying mechanisms of small-molecule AMPK activators on the innate immune and inflammatory responses. (red upward arrows indicate activation/increase and blue downward arrows indicate inhibition/decrease)
Ijms 19 03495 g003
Table 1. Beneficial Effects of AMPK in viral infection.
Table 1. Beneficial Effects of AMPK in viral infection.
PathogenSmall Molecules/ChemicalsAgonist/AntagonistInvolvement of AMPKOutcome (In Vitro/In Vivo)Ref.
Hepatitis C virus (HCV)HCV-HCV infection inhibit AMPKα phosphorylation and Akt-TSC-mTORC1 pathwayAMPK inhibition is required for HCV replication (in vitro)[27,28]
AICAR, Metformin, A769662AgonistRestoration of AMPKα activityAntiviral effects (in vitro)[27,28]
MetforminAgonistType I interferon signaling through AMPK pathway activationInhibits HCV replication (in vitro)[30]
AICARAgonistAMPK activation (Indirect effects counteracted by compound C)Suppression of HCV replication (in vitro)[31]
Hepatitis B virus (HBV)HBV-ROS-dependent AMPK activation in HBV-producing cellsNegatively regulates HBV production[9]
AICAR constitutive active AMPKαAgonistAMPK activation, autophagic flux activationInhibits HBV production (in vitro)[9]
Compound C dominant-negative AMPKαAntagonistAMPK inhibitionEnhances HBV production (in vitro and in vivo)[9]
Vesicular stomatitis virus (VSV)AICARAgonistSTING-dependent signaling activation Type I IFN production and antiviral responses (in vitro)[10]
Compound CAntagonistInhibition of STING-dependent signalingSuppression of IFN-β production (in vitro)[10]
Influenza virusMint3 depletion-AMPK activation Attenuates severe pneumonia by influenza infection (in vivo)[34]
AICARAgonistAMPK activation in Mint3 depletion modelDecreases inflammatory cytokine production in Mint3-deficient macrophages (in vitro)[34]
CurcuminActivatorAMPK activationInhibits influenza A virus infection (in vitro and in vivo)[36]
Human immunodeficiency virus-1Epigallocatechin gallateActivatorAMPK activationAttenuation of Tat-induced human immunodeficiency virus 1 (HIV-1) transactivation[37]
Human adenovirusAdenovirus-Inhibit AMPK activity/signalingVirus induces lipid droplets, presumably associated with obesity (in vitro)[38]
Rift Valley Fever Virus (RVFV)A769662, 2-deoxy-d-glucose (2-DG)AgonistLKB1/AMPK signaling activation; Inhibition of fatty acid synthesisRestriction of viral infection (in vitro)[39]
Herpes simplex virus type 1 (HSV-1)AICAR, Resveratrol, QuercetinActivator/agonistAMPK/Sirt1 activationReduces viral titer and the expression of viral genes (in vitro)[40]
Coxsackievirus B3 (CVB3)--AMPK activation by CVB3Restriction of viral replication; reversed by siRNA against AMPK[41]
AICAR, A769662, MetforminAgonistAMPK activationRestriction of viral replication; improve the survival rate of infected mice (in vitro and in vivo)[41]
Epstein-Barr virus (EBV)LMP1 of EBV-LKB1-AMPK inactivationAMPK inactivation leads to proliferation and transformation of epithelial cells associated with EBV infection (in vitro)[42]
AICARAgonistAMPK activationInhibition of proliferation of nasopharyngeal epithelial cells (in vitro)[42]
Kaposi’s sarcoma-associated herpesvirus (KSHV)AICAR, Metformin, Constitutive active AMPKAgonistAMPK as a KSHV restriction factorInhibits the expression of viral lytic genes and virion production (in vitro)[43]
Compound C, Knockdown of AMPKα1AntagonistAMPK inhibitionEnhances viral lytic gene expression and virion production (in vitro)[43]
Table 2. Detrimental Effects of AMPK in viral infection.
Table 2. Detrimental Effects of AMPK in viral infection.
PathogenSmall Molecules/ChemicalsAgonist/AntagonistInvolvement of AMPKOutcome (In Vitro/In Vivo)Ref.
RotavirusRNAi-AMPK-mediated glycolysis, fatty acid oxidation and autophagyDevelopment of a rotavirus replication-permissive environment (in vitro)[44]
AICAR, Metformin AgonistAMPK activation (AICAR, directly; Metformin, indirectly)Upregulation of the proportion of viral infected cells (in vitro)[44]
DorsomorphinInhibitorInhibition of AMPK activityReduces the number of infected cells (in vitro)[44]
Dengue virusVirus infection-Elevates 3-hydroxy-3-methylglutaryl-CoA reductase activity through AMPK inactivation Promotes the formation of viral replicative complexes (in vitro)[45]
Metformin, A769662 AgonistAMPK activationAntiviral effects (in vitro)[45]
Compound CAntagonistAMPK inhibitionAugments the viral genome copies (in vitro)[45]
Virus infection-AMPK activation; induction of lipophagyIncreases viral replication (in vitro)[46]
Compound C siRNA against AMPKα1AntagonistInhibition of proviral lipophagyDecreases viral replication (in vitro)[46]
Hepatitis B virus (HBV)HBx protein-Decreased ATP, activates AMPK in rat primary hepatocytesAMPK inhibition decreases HBV replication (in vitro)[47]
Compound CAntagonistActivates mTORC1Reduces HBV replication (in vitro)[47]
Ebola virusCompound CAntagonistLess permissive to Ebola virus infection (Similar effects in AMPKα1- or AMPKα2-deleted mouse embryonic fibroblasts)Inhibits EBOV replication in Vero cells (in vitro)[48]
Avian reovirusVirus infection-Upregulates AMPK phosphorylation leading to p38 MAPK activationIncreases virus replication (in vitro)[49]
P17 protein-P17 protein activates AMPK to induce autophagyIncreases virus replication (in vitro)[50]
AICARAgonistAMPK activation (Indirect effects through p38 MAPK)Increases virus replication (in vitro)[49]
Compound CAntagonistAMPK inhibitionDecreases virus replication (in vitro)[49]
Herpes simplex virus type 1 (HSV-1)HSV-1-In early infection, AMPK is down-regulated, and then recovered graduallyAMPK/Sirt1 axis inhibits host apoptosis in early infection (in vitro)[51]
Human immunodeficiency virus-1 (HIV1)Cocaine-Induces AMPK upregulation; AMPK plays a role in energy deficit and metabolic dysfunctionCocaine exposure during HIV infection accelerates neuronal dysfunction (in vitro)[52]
Human cytomegalovirus (HCMV)RNAi-AMPK may activate numerous metabolic pathways during HCMV infectionsiRNA to AMPKα reduces HCMV replication (in vitro)[53,54]
HCMV-Upregulation of host AMPKFavors viral replication (in vitro)[53,54]
Compound CAntagonistInterferes with normal accumulation of viral proteins and alters the core metabolismCompound C inhibits the viral production of HCMV (in vitro); blocks the immediate early phase of viral replication (in vitro)[53,54]
RNAi to AMPK-Blocks glycolytic activation in HCMV-infected cellsRNA-based inhibition of AMPK attenuates HCMV replication (in vitro)[55]
DigitoxinActivatorDigitoxin modulates AMPK-ULK1 and mTOR activity to increase autophagic fluxViral inhibition (in vitro)[56]
Digitoxin + AICAR-Combination reduces autophagyViral replication (in vitro)[56]
HCMV-Induces targeting host protein viperin to mitochondria; viperin is required for AMPK activation and regulate lipid metabolismViperin-dependent lipogenesis promotes viral replication and production by infected host cells (in vitro)[57]
Respiratory syncytial virus (RSV)RSV-RSV induces autophagy through ROS and AMPK activationRSV-induced autophagy favors viral replication (in vitro)[59]
Compound CAntagonistInhibition of AMPK and autophagyCompound C reduces viral gene and protein expression, and total viral titers (in vitro)
Bluetongue virusBluetongue virus-Induces autophagy through activation of AMPKFavors viral replication (in vitro)[60]
Compound C siRNA to AMPKAntagonistInhibits BTV1-induced autophagyAMPK inhibition decreases viral titers (in vitro)
Rabies virusRabies virus Incomplete autophagy induction via CASP2-AMPK-MAPK1/3/11-AKT1-mTOR pathwaysEnhances viral replication (in vitro)[61]
Sendai virusSendai virus-Induces host protein TDRD7, an inhibitor of autophagy-inducing AMPKHost autophagy and viral replication is inhibited by TDRD7 (in vitro)[62]
Compound C, shRNA to AMPKAntagonistInhibition of AMPK activity; inhibits viral proteinAMPK activity is required for viral replication (in vitro)
Snakehead vesiculo-virusSnakehead vesiculo-virus-Downregulates miR-214, which targets AMPKAMPK upregulation promotes viral replication through reduction of IFN-α expression (in vitro)[65]
Table 3. The roles of AMPK in bacterial infection.
Table 3. The roles of AMPK in bacterial infection.
PathogenSmall Molecules/ChemicalsAgonist/AntagonistInvolvement of AMPKOutcome (In Vitro/In Vivo)Ref.
L. pneumophila--Chronic AMPK activation involved in host susceptibility to infection (Direct effects by AMPKα antisense)Bacterial multiplication in host cells with mitochondrial dysfunction[66]
MetforminAgonistBactericidal effects are mediated by mitochondrial ROS production (Indirect)Antimicrobial responses (in vitro and in vivo)[67]
S. typhymurium--S. typhymurium exhibits virulence through lysosomal degradation of SIRT1 and AMPK to impair autophagyBacterial evasion from autophagic clearance (in vitro)[68]
AICARAgonistUpregulation of autophagyIncreased colocalization of salmonella containing vacuole with LC3 (in vitro)[68]
--AMPK activation via TAK1; autophagy initiation by ULK1 phosphorylationAutophagy activation (in vitro)[69]
Compound CAntagonistAMPK inhibitionIncreased bacterial replication by suppression of autophagy (in vitro)[69]
B. abortus--AMPK activation via inositol-requiring enzyme 1 (IRE1)Promote intracellular growth of B. abortus (in vitro)[72]
Compound CAntagonistAMPK inhibition; activation of NADPH oxidase-mediated ROS productionSuppression of intracellular growth (in vitro)[72]
E. coliPiperineAntagonistInhibits ATP-induced pyroptosis by suppressing AMPK activationInhibition of pyroptosis; Attenuation of systemic inflammation (in vitro and in vivo)[73]
ATP MetforminAgonistAMPK activation; increases pyroptosis by inflammasome activationActivation of pyroptosis (in vitro)[73]
B. thuringiensis--AMPK identified by transcriptome and proteome data analysis in vivo (Indirect)Potentially related to regulation of immune defense (Not determined)[74]
H. pylori--TAK1-mediated AMPK activationProtects gastric epithelial cells from H. pylori-induced apoptosis (in vitro)[75]
A-769662 ResveratrolAgonistInhibits H. pylori-induced apoptosis (Direct effects by overexpression of AMPKα)Alleviates H. pylori-induced gastric epithelial cell apoptosis (in vitro)[76]
Compound 13AgonistInhibits H. pylori-induced apoptosis through AMPK-heme oxygenase-1 signalingAlleviates H. pylori-induced gastric epithelial cell apoptosis (in vitro)[76]
Compound CAntagonistInhibitory effects upon compound 13-mediated anti-H. pylori activities (Direct effects by AMPKα1 shRNAs)Aggravates H. pylori-induced gastric epithelial cell apoptosis (in vitro)[76]
P. aeruginosaAICAR MetforminAgonistCounteracts the bacterial effects on the reduction of transepithelial electrical resistance (Indirect effects)Inhibits hyperglycemia-induced bacterial growth; Improve airway epithelial barrier function (in vitro)[77]
S. aureusAICARAgonistAMPK activationReduces bacterial burden and intraocular inflammation; Increases bacterial killing in macrophages (in vitro and in vivo)[78]
Compound CAntagonistDownregulates AMPK activity (Direct effects by AMPKα1 knockout mice)Counteracts AICAR-mediated anti-inflammatory effects (in vivo); Increases susceptibility towards S. aureus endophthalmitis (in vivo)[78]
P. acneEpigallocatechin gallate-Activates AMPK-sterol regulatory element-binding proteins pathway activationAntilipogenic effects in SEB-1 sebocytes (in vitro)[35]
A. baumannii--Activates autophagy through Beclin-1-dependent AMPK/ERK/mTOR pathway (Indirect effects by different A. baumannii strains)Autophagy may promote antimicrobial responses (in vivo)[79]
Mycobacterium tuberculosisMetforminAgonistAMPK activation; Increased mtROS production; Increases phago-lysosomal fusion (Direct effects upon bacterial growth in vitro)Inhibition of intracellular growth of M. tuberculosis (drug-resistant strain; in vitro); Increases the efficacy of conventional TB drugs in vivo [80]
AICARAgonistAMPK-PPARGC1A signaling-mediated autophagy activation; Enhancement of phagosomal maturation (Direct effects by shRNA against AMPKα)Upregulation of antimicrobial responses (in vitro and in vivo)[12]
Compound CAntagonistCounteracts the effects by AICAR upon intracellular inhibition of M. tuberculosis growthDownregulation of antimicrobial responses (in vitro)[12]
Vitamin D (1,25-D3)-Induces autophagy through LL-37 and AMPK activation (Indirect effects upon LL-37 function)Promotes autophagy and antimicrobial response in human monocytes/macrophages (in vitro)[81]
Phenylbutyrate Vitamin D-Induces LL-37-mediated autophagy (Indirect effects; AMPK is involved in LL-37-mediated autophagy)Improves intracellular killing of M. tuberculosis (in vitro)[82]
Gamma-aminobutyric acid (GABA)AgonistInduces autophagy (Direct effects by shRNA against AMPK)Promotes antimicrobial effects against M. tuberculosis (in vitro and in vivo)[83]
Ohmyungsamycins -Activates AMPK and autophagy; Intracellular inhibition of bacterial growth; Amelioration of inflammation (Indirect effects upon host autophagy)Promotes antimicrobial effects against M. tuberculosis (in vitro and in vivo)[26]
Compound CAntagonistBlocks the secretion of neutrophil Matrix metalloproteinase-8 (MMP-8)Neutrophil MMP-8 secretion is related to matrix destruction in human pulmonary TB (in vitro and in human TB lung specimens)[84]
Table 4. The role of AMPK in parasitic infection.
Table 4. The role of AMPK in parasitic infection.
PathogenSmall Molecules/ChemicalsAgonist/AntagonistInvolvement of AMPK (Direct/Indirect)Outcome (In Vitro/In Vivo)Ref.
Hookworm
Nippostrongylus brasiliensis
- AMPKα1 deficiency inhibit IL-13 and CCL17, and defective type 2 immune resistance (Direct effects using by AMPKα1 knockout mice)AMPKα1 suppresses lung injury and drives M2 polarization during infection[89]
L. infantum--Infection leads to a metabolic switch to activate AMPK through the SIRT1-LKB1 axis (Direct effects using by AMPKα1 knockout mice)Ablation of AMPK promotes parasite clearance in vitro and in vivo[90]
S. japonicum Infection-driven IL-7-IL-7R signaling inhibits autophagy;
IL-7 inhibits macrophage autophagy via AMPK
Anti-autophagic IL-7 increases liver pathology (in vivo)[91]
MetforminAgonistDecreases the autophagosome formation in macrophagesin vitro
Compound C siAMPKαAntagonistIncreases autophagosome formation in macrophages (Direct effects by siAMPKα)in vitro
P. falciparum--AMPK activity is suppressed upon infection Decreases Plasmodium hepatic growth[92]
Salicylate Metformin A769662AgonistAMPK activation impairs the intracellular replication of malariaAntimalarial interventions (in vitro and in vivo)
Trypanosoma cruziResveratrol, MetforminAgonistAMPK activation reduces heart oxidative stress (Indirect effects)Reduces heart parasite burden; Protects heart function in Chagas heart disease (in vivo)[93]

Share and Cite

MDPI and ACS Style

Silwal, P.; Kim, J.K.; Yuk, J.-M.; Jo, E.-K. AMP-Activated Protein Kinase and Host Defense against Infection. Int. J. Mol. Sci. 2018, 19, 3495. https://doi.org/10.3390/ijms19113495

AMA Style

Silwal P, Kim JK, Yuk J-M, Jo E-K. AMP-Activated Protein Kinase and Host Defense against Infection. International Journal of Molecular Sciences. 2018; 19(11):3495. https://doi.org/10.3390/ijms19113495

Chicago/Turabian Style

Silwal, Prashanta, Jin Kyung Kim, Jae-Min Yuk, and Eun-Kyeong Jo. 2018. "AMP-Activated Protein Kinase and Host Defense against Infection" International Journal of Molecular Sciences 19, no. 11: 3495. https://doi.org/10.3390/ijms19113495

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop