Next Article in Journal
Photoinduced Antibacterial Activity of the Essential Oils from Eugenia brasiliensis Lam and Piper mosenii C. DC. by Blue Led Light
Next Article in Special Issue
Dissecting the Antimicrobial Composition of Honey
Previous Article in Journal
UPLC-MS-ESI-QTOF Analysis and Antifungal Activity of the Spondias tuberosa Arruda Leaf and Root Hydroalcoholic Extracts
Previous Article in Special Issue
Harnessing the Potential of Killers and Altruists within the Microbial Community: A Possible Alternative to Antibiotic Therapy?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Host-Targeted Therapeutics against Multidrug Resistant Intracellular Staphylococcus aureus

by
Natalia Bravo-Santano
1,
Volker Behrends
1 and
Michal Letek
2,*
1
Health Sciences Research Centre, University of Roehampton, London SW15 4JD, UK
2
Department of Molecular Biology, Area of Microbiology, University of León, 24004 León, Spain
*
Author to whom correspondence should be addressed.
Submission received: 29 October 2019 / Revised: 21 November 2019 / Accepted: 25 November 2019 / Published: 28 November 2019
(This article belongs to the Special Issue Novel Strategies against Pathogenic Bacteria)

Abstract

:
Staphylococcus aureus is a facultative intracellular pathogen that invades and replicates within many types of human cells. S. aureus has shown to rapidly overcome traditional antibiotherapy by developing multidrug resistance. Furthermore, intracellular S. aureus is protected from the last-resort antibiotics—vancomycin, daptomycin, and linezolid—as they are unable to achieve plasma concentrations sufficient for intracellular killing. Therefore, there is an urgent need to develop novel anti-infective therapies against S. aureus infections. Here, we review the current state of the field and highlight the exploitation of host-directed approaches as a promising strategy going forward.

1. Methicillin-Resistant Staphylococcus aureus (MRSA) Methicillin-Resistant Staphylococcus Aureus as an Example of Antibiotic Resistance

Since antibiotics were discovered and introduced into the market, they have become a standard treatment for bacterial infections. Indeed, antibiotics are often administrated as a routine prophylactic measure during medical procedures where bacterial infection may occur (e.g., surgery) [1]. In addition, antibiotics are commonly used in agriculture to prevent bacterial infections and, also, as growth promoters [2]. This increasing exposure to antibiotics has contributed to the emergence of multidrug-resistant bacteria.
In particular, the ability to rapidly acquire antibiotic resistance represents one of the major features of Staphylococcus aureus. In 1959, methicillin was introduced to treat S. aureus infections caused by penicillin-resistant isolates. Two years later, in 1961, the first methicillin-resistant S. aureus isolate was reported in the United Kingdom [3]. Since then, several MRSA clones have been identified over the past decades [4]. In fact, many staphylococcal infections are caused by strains that are resistant to multiple antibiotics, which are associated with higher costs and extended hospitalization periods, as well as higher morbidity and mortality rates [5].
Moreover, although MRSA infections were initially identified as nosocomial infections, the number of MRSA infection cases within healthy community settings has risen in the past two decades, especially in the United States [6,7,8]. Community-acquired MRSA (CA-MRSA) and hospital-acquired MRSA (HA-MRSA) strains differ in their genotypic and phenotypic characteristics. For instance, CA-MRSA isolates are susceptible to most antimicrobials apart from β-lactam antibiotics and erythromycin, whereas HA-MRSA isolates are resistant to most available antibiotics [9].

2. Intracellular MRSA Is Protected from Common Antibiotic Treatments

The opportunistic and facultative intracellular pathogen S. aureus is carried by 30% of the global population [10,11], the anterior nares of the nasal cavity being the most common carriage site [12,13]. During nasal colonization, S. aureus is capable of internalizing into human nasal epithelial cells, and the colonization of the anterior nares increases the risk of developing bacteraemia in persistent carriers. Skin and soft tissue infections are another common portal of entry, which may lead to the colonization of the bloodstream, and, consequently, organ dysfunction and sepsis [14]. Intracellular S. aureus is also associated with recurrent rhinosinusitis, tonsillitis, and chronic osteomyelitis [11]
Host cell invasion and intracellular survival could be used by S. aureus to infect macrophages, spread to secondary points of infection, evade immune recognition, and avoid exposure to last-resort antibiotics [15,16]. Importantly, the serum levels that can be reached without causing toxicity of three last resort antibiotics routinely employed to treat MRSA infection—vancomycin, daptomycin, and linezolid—are not sufficient to achieve intracellular killing and the eradication of this pathogen [15]. As a result, patients are often required to receive long treatments of intravenous vancomycin, which is in stark contrast to the in vitro effective killing of S. aureus observed for this antibiotic [17]. Moreover, clinical infection relapse is not uncommon, suggesting that the intracellular survival of these bacteria facilitates their resistance to the immune system and current antibiotherapies [18]. In fact, antibiotic treatment failure occurs in 20% of patients, leading to an estimated 20,000 deaths per year in the United States alone [19], despite the fact that the clinical isolates often show sensitivity to the administered antibiotics [20].

3. Current Clinical Management of S. aureus Infections

Treatment of S. aureus infections is becoming a real challenge, especially considering the emergence of MRSA strains resistant to last-resort antibiotics (i.e., vancomycin) [21], as well as its protection against current antibiotics once internalized [15].
Clinical management of MRSA infections varies depending on the type of infection, as well as the bacterial strain. Overall, most MRSA infections usually require a prolonged period of antibiotic therapy, and the removal of infected tissue or biomaterial in cases of localized infection or prosthetic joint infections, respectively [22].
The current list of antibiotics available and approved to treat MRSA infections are vancomycin, daptomycin, linezolid, and some other antimicrobials that have been recently developed, including tedizolid, telavancin, oritavancin, dalbavancin, ceftaroline, and ceftobiprole [23]. However, these latter antibiotics are mostly employed to treat skin and soft tissue infections, with vancomycin, daptomycin, and linezolid being the top options for non-topical and/or systemic infections. Nevertheless, these three last-resort antibiotics also present some limitations. For example, daptomycin cannot be used to treat pneumonia [24], whereas linezolid is not suitable for bacteraemia or infective endocarditis treatment [25]. In addition, strains with reduced susceptibility to daptomycin and linezolid, as well as vancomycin-intermediate and vancomycin-resistant S. aureus (VISA and VRSA) strains have already been reported [26]. In addition, and as discussed above, effective doses of vancomycin, daptomycin, or linezolid cannot be achieved intracellularly without causing toxicity to the patient [15].
Furthermore, it is becoming clear that the complex interaction between the pathogen and the host immune system is preventing the development of an efficient vaccine. For example, different staphylococcal virulence factors, such as protein A, possess immunosuppressive effects, which compromise the immunological memory of the host [27]. Indeed, it has been recently demonstrated that vaccination against staphylococcal antigens located on the cell surface may even aggravate the infection [28]. Another important issue regarding the development of a S. aureus vaccine is the use of experimental animals that are not natural hosts of this pathogen [29]. For instance, although some protection against S. aureus can be obtained in mice, the translation of this protective effect into humans has failed repeatedly [27].
Therefore, due to the alarming emergence of antibiotic resistance, the lack of a suitable vaccine against S. aureus infections and the intracellular protection from last-resort antibiotics, novel therapeutics capable of targeting intracellular S. aureus are urgently needed. A recently explored strategy to tackle this problem is the design of antibodies conjugated with bactericidal antibiotics that are specifically released within the host cell [15]. However, in this approach, the antibiotics are only active due to the lower pH of the phagosome, suggesting that this antibody–antibiotic complex may not be effective against cytosolic S. aureus or replicating bacteria within autophagosomes. In addition, the use of antibiotics targeting cell growth or division of the pathogen is still a major selective pressure towards the emergence of resistant strains.

4. Host-Directed Therapies: A Novel Perspective

A recent study has estimated that the development of a new drug that gets market approval costs almost 3 billion United States dollars [30], and this is a process that takes 10–12 years on average [31]. Taking into account the limited lifetime of antibiotics, the consideration of alternative therapeutic strategies aimed at improving the activity of currently available antibiotics might be the only realistic option to combat the antibiotic crisis in the short, medium, and long term [32,33].
As part of their host-defense evasion mechanism, intracellular pathogens—including S. aureus—subvert and exploit a wide range of host factors to support their intracellular survival, targeting multiple pathways to assure their intracellular proliferation [17,34]. Hence, the study of host–pathogen interactions may lead to the identification of potentially novel pathogen-specific drug targets and/or host-directed therapeutics [17,35].
Repurposing commercially available drugs that target host pathways hijacked by intracellular pathogens is a particularly interesting strategy [33]. The main advantage of using repurposed drugs is that they show minimal toxicity to the host cell, and that they have already been approved for other clinical purposes, which would significantly reduce the necessary time to have these drugs in the market [34].
In addition, the use of repurposed host-targeted therapeutics in combination with traditional antibiotherapy may increase the lifetime of currently available antibiotics by preventing intracellular pathogens’ escape from antibacterial treatment within host cells. This may shorten the length of the treatment required for chronic infections and it should lower the risk of infection relapse.
Indeed, the research focused on host-directed therapies has already identified many drugs that have shown positive effects against intracellular bacterial pathogens (Table 1). These examples provide proof-of-principle about the potential of this approach to improve the outcome of bacterial infections caused by these pathogens. Such therapies have shown the ability to achieve intracellular killing, and they also possess the benefit of preventing the development of antimicrobial resistance [36,37].
Host-directed approaches require a deep understanding of the host–pathogen interactions. In order to gain novel insights about such interactions, individual or combined strategies that include-omics approaches (genomics, transcriptomics, proteomics, and metabolomics) and computational biology have shown great promise. Certainly, genome-wide screens employing RNA interference (RNAi), microarray analyses, and chemical libraries have already been employed to identify host genes required for bacterial entry and intracellular growth of intracellular pathogens such as MRSA [35].
RNAi approaches allow loss-of-function phenotypic analyses to discover novel host-directed targets. For instance, genome-wide RNA interference assays have been performed in Drosophila cells to identify host factors exploited and subverted by two intracellular pathogens: Chlamydia caviae and Listeria monocytogenes. In C. caviae infections, the depletion of human genes TOMM40 or TOMM22, which are involved in protein transport to the mitochondria, resulted in a reduction of intracellular bacteria within mammalian cells [79]. Moreover, 116 host genes were identified to be important during L. monocytogenes infection and implicated in entry, phagosomal escape, and intracellular replication [80].
Interestingly, a study that employed a genetic approach combining RNAi and automated microscopy revealed the importance of serine-threonine protein kinases (Akt1) in the intracellular survival of several intracellular pathogens. Akt1 kinase acts as a key regulator of several guanosine triphosphatases (GTPases), which are involved in essential host pathways exploited by intracellular pathogens. For example, Salmonella Typhimurium is able to activate host Akt1 to control actin dynamics via p21(Cdkn1a)-activated kinase 4 (PAK4) and, hence, facilitate its intracellular survival [50]. Moreover, we recently screened for novel host targets that are essential for the host cell infection of MRSA by means of functional genomics. More specifically, we employed an unbiased short-hairpin RNA (shRNA) lentiviral library to screen 16,000 human genes during intracellular S. aureus infection, identifying several host genes important for intracellular MRSA [81]. In particular, we found that silencing the human gene TRAM2 resulted in a significant reduction of intracellular MRSA, whereas host cell viability was restored, showing its importance during intracellular infection. TRAM2 is an interactive partner of the sarco/endoplasmic reticulum Ca2+ ATPase (SERCA) [82]. Accordingly, we found that very low doses of thapsigargin (an inhibitor of the SERCA pumps) could be used to stop S. aureus intracellular survival in combination with conventional antibiotics [81].
The use of metabolomics also represents an interesting approach to gain a better understanding of the metabolic scenario upon intracellular bacterial infection [83,84]. It has been shown that several intracellular pathogens trigger host cell metabolism changes to support its intracellular survival [85]. For example, Legionella pneumophila induce a Warburg-like effect in the host cell by interacting with mitochondria to favour its own replication [86]. Shigella flexneri re-routes host central carbon metabolism to obtain an abundant nutrient-flux through the glycolytic pathway that allows its intracellular survival [87]. Moreover, host cholesterol import is required by intracellular Mycobacterium tuberculosis to persist inside both macrophages and mice lungs [88]. On the other hand, a reduction in nutrient uptake as well as nucleotide biosynthesis has been recently observed in human airway epithelial cells infected with S. aureus [89].
We recently characterized host cell physiology of MRSA-infected cells using mass spectrometry-based metabolomics [17,90]. We found that S. aureus infection leads to starvation-induced autophagy due to a block in central carbon metabolism, which is mediated by the energy sensor AMPK (AMP-activated protein kinase). Consequently, a treatment with the AMPK inhibitor dorsomorphin halted intracellular S. aureus in HeLa and Human Umbilical Vein Endothelial Cells (HUVECs) [17]. Accordingly, treatments with autophagy inhibitors protect mice from MRSA pneumonia [91].
Once specific host factors or metabolic pathways required by the intracellular pathogen have been described, blocking these pathways—for example, by using drug inhibitors—may be considered as a novel strategy to control bacterial infections. For example, Akt1 and protein kinase A (PKA) inhibitors have been shown to reduce intracellular load of M. tuberculosis and S. aureus Typhimurium in infected human macrophages [50]. Moreover, host-directed therapies could also aim to enhance host cellular responses against intracellular pathogens to activate innate and adaptive host immune responses or to modulate disproportionate inflammation. Hence, host-directed therapies also include immunomodulatory agents—including monoclonal antibodies or nutritional products—as well as cellular therapy (Table 1). For example, monoclonal antibodies anti-TNFα and anti-interleukins, which reduce tissue-destructive inflammation by cytokine neutralisation, have shown a bactericidal effect against the intracellular pathogens M. tuberculosis and Helicobacter pylori [38,39,92]. Further, the nutritional product vitamin D3 is also effective against these two pathogens, and its mechanism of action has been described as an activation and enhancement of host antimicrobial defences [78,93].
In addition, several studies have already shown that existing and approved drugs, which are used for other clinical purposes unrelated to infection biology, can influence the outcome of bacterial infections. The main advantage of finding “repurposed drugs” is that they are already clinically approved, shortening the time to reach the clinic [35,53]. An example of repurposed drug is the histone deacetylase inhibitor sulforaphane, which induces the expression of leukocyte protease inhibitor and β-defensin 2 and, therefore, increases the activity of antibiotics against the multidrug-resistant strain Neisseria gonorrhoeae [70]. Similarly, imatinib—an ABL tyrosine kinase inhibitor, commercialized as Gleevec—is under early clinical trials to treat complicated infections caused by M. tuberculosis [57,94].
To further identify other host-targeted therapeutics against intracellular MRSA, we recently screened a panel of 133 FDA-approved drugs with known host targets to test whether blocking these targets had an effect on intracellular bacterial survival [34]. Interestingly, we found that ibrutinib, a host kinase inhibitor, significantly increased host cell viability and reduced bacterial survival at clinically relevant concentrations. We also determined the mechanism of action of ibrutinib by mass spectrometry-based phosphoproteomics, finding very promising host targets, such as Ephrin receptor 2 (EPHA2), that could be used to further develop a more specific therapy against intracellular S. aureus [34].
In summary, Table 2 compresses the host factors that have been identified as important for intracellular S. aureus internalization, survival, and/or replication. It also includes the putative function of each pathway and the potential host-directed drug (if known) to tackle such pathway with the aim to impair intracellular S. aureus.

5. Conclusions

Future anti-infective therapies may consist in the combination of conventional antibiotics targeting bacterial survival outside of the host cell with host-directed therapies to efficiently eliminate intracellular pathogens. Therefore, a better understanding of host–S. aureus interactions during intracellular infection may lead to novel therapeutic strategies by targeting those host cellular pathways exploited by this versatile pathogen.

Author Contributions

Conceptualization, V.B. and M.L.; writing—original draft preparation, N.B.-S.; writing—review and editing, N.B.-S., V.B., and M.L.; supervision, V.B. and M.L.; project administration, V.B. and M.L.; funding acquisition, V.B. and M.L.

Funding

This work was supported by a Roehampton Vice Chancellor’s Scholarship to N.B.-S. and intramural funding from the University of Roehampton to V.B. and M.L.

Acknowledgments

We thank Jolanta Opacka-Juffry, Cokro Leksmono, and Martha Villegas-Montes for their support at the University of Roehampton during the completion of our research work cited in this manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Agodi, A.; Barchitta, M.; Maugeri, A.; Sodano, L.; Pasquarella, C. Appropriate perioperative antibiotic prophylaxis: Challenges, strategies, and quality indicators. Epidemiol. Prev. 2015, 39, 27–32. [Google Scholar]
  2. McEwen, S.A.; Collignon, P.J. Antimicrobial Resistance: A One Health Perspective. Microbiol. Spectr. 2018, 6. [Google Scholar] [CrossRef] [PubMed]
  3. Jevons, M.P. “Celbenin”—Resistant Staphylococci. Br. Med. J. 1961, 1, 124–125. [Google Scholar] [CrossRef]
  4. Hiramatsu, K.; Cui, L.; Kuroda, M.; Ito, T. The emergence and evolution of methicillin-resistant Staphylococcus aureus. Trends Microbiol. 2001, 9, 486–493. [Google Scholar] [CrossRef]
  5. Ippolito, G.; Leone, S.; Lauria, F.N.; Nicastri, E.; Wenzel, R.P. Methicillin-resistant Staphylococcus aureus: The superbug. Int. J. Infect. Dis. 2010, 14, 7–11. [Google Scholar] [CrossRef] [PubMed]
  6. Drews, T.D.; Temte, J.L.; Fox, B.C. Community-associated methicillin-resistant Staphylococcus aureus: Review of an emerging public health concern. Wis. Med. J. 2006, 105, 52–57. [Google Scholar]
  7. David, M.Z.; Daum, R.S. Community-associated methicillin-resistant Staphylococcus aureus: Epidemiology and clinical consequences of an emerging epidemic. Clin. Microbiol. Rev. 2010, 23, 616–687. [Google Scholar] [CrossRef]
  8. Otto, M. Community-associated MRSA: What makes them special? Int. J. Med. Microbiol. 2013, 303, 324–330. [Google Scholar] [CrossRef]
  9. Naimi, T.S.; Ledell, K.H.; Como-sabetti, K.; Borchardt, S.M.; Boxrud, D.J.; Johnson, S.K.; Fridkin, S.; Boyle, C.O.; Danila, R.N.; Lynfield, R. Comparison of Community- and Health Care-Associated Methicillin-Staphylococcus aureus Infection. JAMA 2003, 290, 2976–2984. [Google Scholar] [CrossRef]
  10. Sollid, J.U.E.; Furberg, A.S.; Hanssen, A.M.; Johannessen, M. Staphylococcus aureus: Determinants of human carriage. Infect. Genet. Evol. 2014, 21, 531–541. [Google Scholar] [CrossRef]
  11. Sakr, A.; Brégeon, F.; Mege, J.-L.; Rolain, J.-M.; Blin, O. Staphylococcus aureus nasal colonization: An update on mechanisms, epidemiology, risk factors and subsequent infections. Front. Microbiol. 2018, 9, 1–15. [Google Scholar] [CrossRef] [PubMed]
  12. François, P.; Vaudaux, P.; Foster, T.J.; Lew, D.P. Host-Bacteria Interactions in Foreign Body Infections. Infect. Control Hosp. Epidemiol. 1996, 17, 514–520. [Google Scholar] [CrossRef] [PubMed]
  13. Wertheim, H.F.; Melles, D.C.; Vos, M.C.; van Leeuwen, W.; van Belkum, A.; Verbrugh, H.A.; Nouwen, J.L. The role of nasal carriage in Staphylococcus aureus infections. Lancet Infect. Dis. 2005, 5, 751–762. [Google Scholar] [CrossRef]
  14. Yoshikawa, T.T.; Strausbaugh, L.J. Methicillin-resistant Staphylococcus aureus. In Infection Management for Geriatrics in Long-Term Care Facilities, 2nd ed.; CRC Press: Boca Raton, FL, USA, 2006; ISBN 9781420021110. [Google Scholar]
  15. Lehar, S.M.; Pillow, T.; Xu, M.; Staben, L.; Kajihara, K.K.; Vandlen, R.; DePalatis, L.; Raab, H.; Hazenbos, W.L.; Hiroshi Morisaki, J.; et al. Novel antibody–antibiotic conjugate eliminates intracellular S. aureus. Nature 2015, 527, 323–328. [Google Scholar] [CrossRef] [PubMed]
  16. Jorch, S.K.; Surewaard, B.G.; Hossain, M.; Peiseler, M.; Deppermann, C.; Deng, J.; Bogoslowski, A.; van der Wal, F.; Omri, A.; Hickey, M.J.; et al. Peritoneal GATA6+ macrophages function as a portal for Staphylococcus aureus dissemination. J. Clin. Investig. 2019. [Google Scholar] [CrossRef]
  17. Bravo-Santano, N.; Ellis, J.K.; Mateos, L.M.; Calle, Y.; Keun, H.C.; Behrends, V.; Letek, M. Intracellular Staphylococcus aureus Modulates Host Central Carbon Metabolism To Activate Autophagy. mSphere 2018, 3, e00374-18. [Google Scholar] [CrossRef]
  18. Kullar, R.; Davis, S.L.; Levine, D.P.; Rybak, M.J. Impact of vancomycin exposure on outcomes in patients with methicillin-resistant Staphylococcus aureus bacteremia: Support for consensus guidelines suggested targets. Clin. Infect. Dis. 2011, 52, 975–981. [Google Scholar] [CrossRef]
  19. Kourtis, A.P.; Hatfield, K.; Baggs, J.; Mu, Y.; See, I.; Epson, E.; Nadle, J.; Kainer, M.A.; Dumyati, G.; Petit, S.; et al. Vital signs: Epidemiology and recent trends in methicillin-resistant and in methicillin-susceptible Staphylococcus aureus bloodstream infections—United States. Morb. Mortal. Wkly. Rep. 2019, 68, 214. [Google Scholar] [CrossRef]
  20. Walraven, C.J.; North, M.S.; Marr-Lyon, L.; Deming, P.; Sakoulas, G.; Mercier, R.C. Site of infection rather than vancomycin MIC predicts vancomycin treatment failure in methicillin-resistant Staphylococcus aureus bacteraemia. J. Antimicrob. Chemother. 2011, 66, 2386–2392. [Google Scholar] [CrossRef]
  21. Howden, B.P.; Davies, J.K.; Johnson, P.D.R.; Stinear, T.P.; Grayson, M.L. Reduced Vancomycin Susceptibility in Staphylococcus aureus, Including Vancomycin-Intermediate and Heterogeneous Vancomycin-Intermediate Strains: Resistance Mechanisms, Laboratory Detection, and Clinical Implications. Clin. Microbiol. Rev. 2010, 23, 99–139. [Google Scholar] [CrossRef]
  22. Tong, S.Y.C.; Davis, J.S.; Eichenberger, E.; Holland, T.L.; Fowler, V.G. Staphylococcus aureus infections: Epidemiology, pathophysiology, clinical manifestations, and management. Clin. Microbiol. Rev. 2015, 28, 603–661. [Google Scholar] [CrossRef] [PubMed]
  23. Boswihi, S.S.; Udo, E.E. Methicillin-resistant Staphylococcus aureus: An update on the epidemiology, treatment options and infection control. Curr. Med. Res. Pract. 2018, 8, 18–24. [Google Scholar] [CrossRef]
  24. Raja, A.; LaBonte, J.; Lebbos, J.; Kirkpatrick, P. Daptomycin. Nat. Rev. Drug Discov. 2003, 2, 943. [Google Scholar] [CrossRef] [PubMed]
  25. Watkins, R.R.; Lemonovich, T.L.; File, T.M., Jr. An evidence-based review of linezolid for the treatment of methicillin-resistant Staphylococcus aureus (MRSA): Place in therapy. Core Evid. 2012, 7, 131–143. [Google Scholar] [CrossRef]
  26. Nannini, E.; Murray, B.E.; Arias, C.A. Resistance or decreased susceptibility to glycopeptides, daptomycin, and linezolid in methicillin-resistant Staphylococcus aureus. Curr. Opin. Pharm. 2010, 10, 516–521. [Google Scholar] [CrossRef]
  27. Kobayashi, S.D.; Deleo, F.R. Staphylococcus aureus Protein A Promotes Immune Suppression. MBio 2013, 4, e00746-13. [Google Scholar] [CrossRef]
  28. Spaulding, A.R.; Salgado-Pabón, W.; Merriman, J.A.; Stach, C.S.; Ji, Y.; Gillman, A.N.; Peterson, M.L.; Schlievert, P.M. Vaccination against Staphylococcus aureus pneumonia. J. Infect. Dis. 2014, 209, 1955–1962. [Google Scholar] [CrossRef]
  29. Brown, A.F.; Leech, J.M.; Rogers, T.R.; McLoughlin, R.M. Staphylococcus aureus colonization: Modulation of host immune response and impact on human vaccine design. Front. Immunol. 2014, 4, 507. [Google Scholar] [CrossRef]
  30. DiMasi, J.A.; Grabowski, H.G.; Hansen, R.W. Innovation in the pharmaceutical industry: New estimates of R&D costs. J. Health Econ. 2016, 47, 20–33. [Google Scholar]
  31. Sun, W.; Sanderson, P.E.; Zheng, W. Drug combination therapy increases successful drug repositioning. Drug Discov. Today 2016, 21, 1189–1195. [Google Scholar] [CrossRef]
  32. Zheng, W.; Sun, W.; Simeonov, A. Drug repurposing screens and synergistic drug-combinations for infectious diseases. Br. J. Pharm. 2018, 175, 181–191. [Google Scholar] [CrossRef] [PubMed]
  33. Chiang, C.Y.; Uzoma, I.; Moore, R.T.; Gilbert, M.; Duplantier, A.J.; Panchal, R.G. Mitigating the impact of antibacterial drug resistance through host-directed therapies: Current progress, outlook, and challenges. MBio 2018, 9, e1092-17. [Google Scholar] [CrossRef] [PubMed]
  34. Bravo-Santano, N.; Stölting, H.; Cooper, F.; Bileckaja, N.; Majstorovic, A.; Ihle, N.; Mateos, L.M.; Calle, Y.; Behrends, V.; Letek, M. Host-directed kinase inhibitors act as novel therapies against intracellular Staphylococcus aureus. Sci. Rep. 2019, 9, 4876. [Google Scholar] [CrossRef] [PubMed]
  35. Schwegmann, A.; Brombacher, F. Host-directed drug targeting of factors hijacked by pathogens. Sci. Signal. 2008, 1, re8. [Google Scholar] [CrossRef]
  36. Zumla, A.; Rao, M.; Wallis, R.S.; Kaufmann, S.H.E.; Rustomjee, R.; Mwaba, P.; Vilaplana, C.; Yeboah-Manu, D.; Chakaya, J.; Ippolito, G.; et al. Host-directed therapies for infectious diseases: Current status, recent progress, and future prospects. Lancet Infect. Dis. 2016, 16, 47–63. [Google Scholar] [CrossRef] [Green Version]
  37. Kaufmann, S.H.E.; Dorhoi, A.; Hotchkiss, R.S.; Bartenschlager, R. Host-directed therapies for bacterial and viral infections. Nat. Rev. Drug Discov. 2018, 17, 35–56. [Google Scholar] [CrossRef]
  38. Wallis, R.S.; van Vuuren, C.; Potgieter, S. Adalimumab Treatment of Life-Threatening Tuberculosis. Clin. Infect. Dis. 2009, 48, 1429–1432. [Google Scholar] [CrossRef] [Green Version]
  39. Wroblewski, L.E.; Peek, R.M.; Wilson, K.T. Helicobacter pylori and Gastric Cancer: Factors That Modulate Disease Risk. Clin. Microbiol. Rev. 2010, 23, 713–739. [Google Scholar] [CrossRef] [Green Version]
  40. Scanlon, K.M.; Skerry, C.; Carbonetti, N.H. Novel therapies for the treatment of pertussis disease. Pathog. Dis. 2015, 73, ftv074. [Google Scholar] [CrossRef] [Green Version]
  41. Datta, M.; Via, L.E.; Kamoun, W.S.; Liu, C.; Chen, W.; Seano, G.; Weiner, D.M.; Schimel, D.; England, K.; Martin, J.D.; et al. Anti-vascular endothelial growth factor treatment normalizes tuberculosis granuloma vasculature and improves small molecule delivery. Proc. Natl. Acad. Sci. USA 2015, 112, 1827–1832. [Google Scholar] [CrossRef] [Green Version]
  42. Oehlers, S.H.; Cronan, M.R.; Scott, N.R.; Thomas, M.I.; Okuda, K.S.; Walton, E.M.; Beerman, R.W.; Crosier, P.S.; Tobin, D.M. Interception of host angiogenic signalling limits mycobacterial growth. Nature 2014, 517, 612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Singh, A.; Mohan, A.; Dey, A.B.; Mitra, D.K. Inhibiting the Programmed Death 1 Pathway Rescues Mycobacterium tuberculosis–Specific Interferon γ—Producing T Cells From Apoptosis in Patients with Pulmonary Tuberculosis. J. Infect. Dis. 2013, 208, 603–615. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Jurado, J.O.; Alvarez, I.B.; Pasquinelli, V.; Martínez, G.J.; Quiroga, M.F.; Abbate, E.; Musella, R.M.; Chuluyan, H.E.; García, V.E. Programmed Death (PD)-1: PD-Ligand 1/PD-Ligand 2 Pathway Inhibits T Cell Effector Functions during Human Tuberculosis. J. Immunol. 2008, 181, 116–125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Ravimohan, S.; Tamuhla, N.; Steenhoff, A.P.; Letlhogile, R.; Nfanyana, K.; Bellamy, S.L.; MacGregor, R.R.; Gross, R.; Weissman, D.; Bisson, G.P. Immunological profiling of tuberculosis-associated immune reconstitution inflammatory syndrome and non-immune reconstitution inflammatory syndrome death in HIV-infected adults with pulmonary tuberculosis starting antiretroviral therapy: A prospective obse. Lancet Infect. Dis. 2015, 15, 429–438. [Google Scholar] [CrossRef] [Green Version]
  46. Tobin, D.M.; Roca, F.J.; Ray, J.P.; Ko, D.C.; Ramakrishnan, L. An Enzyme That Inactivates the Inflammatory Mediator Leukotriene B4 Restricts Mycobacterial Infection. PLoS ONE 2013, 8, e67828. [Google Scholar] [CrossRef] [Green Version]
  47. Martins, M.; Bleiss, W.; Marko, A.; Ordway, D.; Viveiros, M.; Leandro, C.; Pacheco, T.; Molnar, J.; Kristiansen, J.E.; Amaral, L. Clinical concentrations of thioridazine enhance the killing of intracellular methicillin-resistant Staphylococcus aureus: An in vivo, ex vivo and electron microscopy study. In Vivo 2004, 18, 787–794. [Google Scholar]
  48. Rikihisa, Y.; Zhang, Y.; Park, J. Role of Ca2+ and calmodulin in ehrlichial infection in macrophages. Infect. Immun. 1995, 63, 2310–2316. [Google Scholar]
  49. Amaral, L.; Kristiansen, J.E.; Frølund Thomsen, V.; Markovich, B. The effects of chlorpromazine on the outer cell wall of Salmonella Typhimurium in ensuring resistance to the drug. Int. J. Antimicrob. Agents 2000, 14, 225–229. [Google Scholar] [CrossRef]
  50. Kuijl, C.; Savage, N.D.L.; Marsman, M.; Tuin, A.W.; Janssen, L.; Egan, D.A.; Ketema, M.; van den Nieuwendijk, R.; van den Eeden, S.J.F.; Geluk, A.; et al. Intracellular bacterial growth is controlled by a kinase network around PKB/AKT1. Nature 2007, 450, 725. [Google Scholar] [CrossRef]
  51. Skerry, C.; Scanlon, K.; Rosen, H.; Carbonetti, N.H. Sphingosine-1-phosphate Receptor Agonism Reduces Bordetella pertussis—Mediated Lung Pathology. J. Infect. Dis. 2015, 211, 1883–1886. [Google Scholar] [CrossRef] [Green Version]
  52. Koh, G.C.K.W.; Maude, R.R.; Schreiber, M.F.; Limmathurotsakul, D.; Wiersinga, W.J.; Wuthiekanun, V.; Lee, S.J.; Mahavanakul, W.; Chaowagul, W.; Chierakul, W.; et al. Glyburide Is Anti-inflammatory and Associated with Reduced Mortality in Melioidosis. Clin. Infect. Dis. 2011, 52, 717–725. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Czyz, D.M.; Potluri, L.-P.; Jain-Gupta, N.; Riley, S.P.; Martinez, J.J.; Steck, T.L.; Crosson, S.; Shuman, H.A.; Gabay, J.E. Host-directed antimicrobial drugs with broad-spectrum efficacy against intracellular bacterial pathogens. MBio 2014, 5, e01534-14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Bernard, G.R.; Wheeler, A.P.; Russell, J.A.; Schein, R.; Summer, W.R.; Steinberg, K.P.; Fulkerson, W.J.; Wright, P.E.; Christman, B.W.; Dupont, W.D.; et al. The Effects of Ibuprofen on the Physiology and Survival of Patients with Sepsis. N. Engl. J. Med. 1997, 336, 912–918. [Google Scholar] [CrossRef] [PubMed]
  55. Vilaplana, C.; Marzo, E.; Tapia, G.; Diaz, J.; Garcia, V.; Cardona, P.-J. Ibuprofen Therapy Resulted in Significantly Decreased Tissue Bacillary Loads and Increased Survival in a New Murine Experimental Model of Active Tuberculosis. J. Infect. Dis. 2013, 208, 199–202. [Google Scholar] [CrossRef] [Green Version]
  56. Ivanyi, J.; Zumla, A. Nonsteroidal Antiinflammatory Drugs for Adjunctive Tuberculosis Treatment. J. Infect. Dis. 2013, 208, 185–188. [Google Scholar] [CrossRef]
  57. Napier, R.J.; Rafi, W.; Cheruvu, M.; Powell, K.R.; Zaunbrecher, M.A.; Bornmann, W.; Salgame, P.; Shinnick, T.M.; Kalman, D. Imatinib-sensitive tyrosine kinases regulate mycobacterial pathogenesis and represent therapeutic targets against tuberculosis. Cell Host Microbe 2011, 10, 475–485. [Google Scholar] [CrossRef] [Green Version]
  58. Lin, M.; Den Dulk-Ras, A.; Hooykaas, P.J.J.; Rikihisa, Y. Anaplasma phagocytophilum AnkA secreted by type IV secretion system is tyrosine phosphorylated by Abl-1 to facilitate infection. Cell. Microbiol. 2007, 9, 2644–2657. [Google Scholar] [CrossRef]
  59. Bruce Light, R. Indomethacin and Acetylsalicylic Acid Reduce Intrapulmonary Shunt in Experimental Pneumococcal Pneumonia. Am. Rev. Respir. Dis. 1986, 134, 520–525. [Google Scholar]
  60. Singhal, A.; Jie, L.; Kumar, P.; Hong, G.S.; Leow, M.K.-S.; 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]
  61. Pirinen, E.; Cantó, C.; Jo, Y.S.; Morato, L.; Zhang, H.; Menzies, K.J.; Williams, E.G.; Mouchiroud, L.; Moullan, N.; Hagberg, C.; et al. Pharmacological Inhibition of Poly(ADP-Ribose) Polymerases Improves Fitness and Mitochondrial Function in Skeletal Muscle. Cell Metab. 2014, 19, 1034–1041. [Google Scholar] [CrossRef] [Green Version]
  62. Coussens, A.K.; Wilkinson, R.J.; Martineau, A.R. Phenylbutyrate Is Bacteriostatic against Mycobacterium tuberculosis and Regulates the Macrophage Response to Infection, Synergistically with 25-Hydroxy-Vitamin D3. PLoS Pathog. 2015, 11, e1005007. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Lieberman, L.A.; Higgins, D.E. A small-molecule screen identifies the antipsychotic drug pimozide as an inhibitor of Listeria monocytogenes infection. Antimicrob. Agents Chemother. 2009, 53, 756–764. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Blum, C.A.; Nigro, N.; Briel, M.; Schuetz, P.; Ullmer, E.; Suter-Widmer, I.; Winzeler, B.; Bingisser, R.; Elsaesser, H.; Drozdov, D.; et al. Adjunct prednisone therapy for patients with community-acquired pneumonia: A multicentre, double-blind, randomised, placebo-controlled trial. Lancet 2015, 385, 1511–1518. [Google Scholar] [CrossRef]
  65. Critchley, J.A.; Young, F.; Orton, L.; Garner, P. Corticosteroids for prevention of mortality in people with tuberculosis: A systematic review and meta-analysis. Lancet Infect. Dis. 2013, 13, 223–237. [Google Scholar] [CrossRef]
  66. Ho Sui, S.J.; Lo, R.; Fernandes, A.R.; Caulfield, M.D.G.; Lerman, J.A.; Xie, L.; Bourne, P.E.; Baillie, D.L.; Brinkman, F.S.L. Raloxifene attenuates Pseudomonas aeruginosa pyocyanin production and virulence. Int. J. Antimicrob. Agents 2012, 40, 246–251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Mortensen, E.M.; Pugh, M.J.; Copeland, L.A.; Restrepo, M.I.; Cornell, J.E.; Anzueto, A.; Pugh, J.A. Impact of statins and angiotensin-converting enzyme inhibitors on mortality of subjects hospitalised with pneumonia. Eur. Respir. J. 2008, 31, 611–617. [Google Scholar] [CrossRef] [Green Version]
  68. Chalmers, J.D.; Singanayagam, A.; Murray, M.P.; Hill, A.T. Prior Statin Use Is Associated with Improved Outcomes in Community-acquired Pneumonia. Am. J. Med. 2008, 121, 1002–1007. [Google Scholar] [CrossRef]
  69. Parihar, S.P.; Guler, R.; Khutlang, R.; Lang, D.M.; Hurdayal, R.; Mhlanga, M.M.; Suzuki, H.; Marais, A.D.; Brombacher, F. Statin Therapy Reduces the Mycobacterium tuberculosis Burden in Human Macrophages and in Mice by Enhancing Autophagy and Phagosome Maturation. J. Infect. Dis. 2014, 209, 754–763. [Google Scholar] [CrossRef] [Green Version]
  70. Yedery, D.R.; Jerse, E.A. Augmentation of Cationic Antimicrobial Peptide Production with Histone Deacetylase Inhibitors as a Novel Epigenetic Therapy for Bacterial Infections. Antibiotics 2015, 4, 44–61. [Google Scholar] [CrossRef]
  71. Lieberman, L.A.; Higgins, D.E. Inhibition of Listeria monocytogenes infection by neurological drugs. Int. J. Antimicrob. Agents 2010, 35, 292–296. [Google Scholar] [CrossRef] [Green Version]
  72. Amaral, L.; Viveiros, M. Why thioridazine in combination with antibiotics cures extensively drug-resistant Mycobacterium tuberculosis infections. Int. J. Antimicrob. Agents 2012, 39, 376–380. [Google Scholar] [CrossRef] [PubMed]
  73. Martins, M.; Viveiros, M.; Amaral, L. Inhibitors of Ca2+ and K+ Transport Enhance Intracellular Killing of M. tuberculosis by Non-killing Macrophages. In Vivo 2008, 22, 69–75. [Google Scholar] [PubMed]
  74. Gupta, S.; Tyagi, S.; Bishai, W.R. Verapamil increases the bactericidal activity of bedaquiline against Mycobacterium tuberculosis in a mouse model. Antimicrob. Agents Chemother. 2015, 59, 673–676. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Salin, O.P.; Pohjala, L.L.; Saikku, P.; Vuorela, H.J.; Leinonen, M.; Vuorela, P.M. Effects of coadministration of natural polyphenols with doxycycline or calcium modulators on acute Chlamydia pneumoniae infection in vitro. J. Antibiot. 2011, 64, 747. [Google Scholar] [CrossRef] [Green Version]
  76. Rao, M.; Valentini, D.; Zumla, A.; Maeurer, M. Evaluation of the efficacy of valproic acid and suberoylanilide hydroxamic acid (vorinostat) in enhancing the effects of first-line tuberculosis drugs against intracellular Mycobacterium tuberculosis. Int. J. Infect. Dis. 2018, 69, 78–84. [Google Scholar] [CrossRef] [Green Version]
  77. Mayer-Barber, K.D.; Andrade, B.B.; Oland, S.D.; Amaral, E.P.; Barber, D.L.; Gonzales, J.; Derrick, S.C.; Shi, R.; Kumar, N.P.; Wei, W.; et al. Host-directed therapy of tuberculosis based on interleukin-1 and type I interferon crosstalk. Nature 2014, 511, 99. [Google Scholar] [CrossRef] [Green Version]
  78. Guo, L.; Chen, W.; Zhu, H.; Chen, Y.; Wan, X.; Yang, N.; Xu, S.; Yu, C.; Chen, L. Helicobacter pylori Induces Increased Expression of the Vitamin D Receptor in Immune Responses. Helicobacter 2013, 19, 37–47. [Google Scholar] [CrossRef]
  79. Derré, I.; Pypaert, M.; Dautry-Varsat, A.; Agaisse, H. RNAi Screen in Drosophila Cells Reveals the Involvement of the Tom Complex in Chlamydia Infection. PLoS Pathog. 2007, 3, e155. [Google Scholar] [CrossRef] [Green Version]
  80. Cheng, L.W.; Viala, J.P.M.; Stuurman, N.; Wiedemann, U.; Vale, R.D.; Portnoy, D.A. Use of RNA interference in Drosophila S2 cells to identify host pathways controlling compartmentalization of an intracellular pathogen. Proc. Natl. Acad. Sci. USA 2005, 102, 13646–13651. [Google Scholar] [CrossRef] [Green Version]
  81. Bravo-Santano, N.; Capilla-Lasheras, P.; Mateos, L.M.; Calle, Y.; Behrends, V.L.M. Identification of novel targets for host-directed therapeutics against intracellular Staphylococcus aureus. Sci. Rep. 2019, 9, 1–12. [Google Scholar] [CrossRef] [Green Version]
  82. Stefanovic, B.; Stefanovic, L.; Schnabl, B.; Bataller, R.; Brenner, D.A. TRAM2 protein interacts with endoplasmic reticulum Ca2+ pump Serca2b and is necessary for collagen type I synthesis. Mol. Cell. Biol. 2004, 24, 1758–1768. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Kint, G.; Fierro, C.; Marchal, K.; Vanderleyden, J.; De Keersmaecker, S.C.J. Integration of ‘omics’ data: Does it lead to new insights into host–microbe interactions? Future Microbiol. 2010, 5, 313–328. [Google Scholar] [CrossRef] [PubMed]
  84. Jean Beltran, P.M.; Federspiel, J.D.; Sheng, X.; Cristea, I.M. Proteomics and integrative omic approaches for understanding host-pathogen interactions and infectious diseases. Mol. Syst. Biol. 2017, 13. [Google Scholar] [CrossRef] [PubMed]
  85. Eisenreich, W.; Heesemann, J.; Rudel, T.; Goebel, W. Metabolic host responses to infection by intracellular bacterial pathogens. Front. Cell. Infect. Microbiol. 2013, 3, 24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Escoll, P.; Song, O.R.O.-R.; Viana, F.; Steiner, B.; Lagache, T.; Olivo-Marin, J.-C.J.C.; Impens, F.; Brodin, P.; Hilbi, H.; Buchrieser, C. Legionella pneumophila Modulates Mitochondrial Dynamics to Trigger Metabolic Repurposing of Infected Macrophages. Cell Host Microbe 2017, 22, 302–316. [Google Scholar] [CrossRef] [PubMed]
  87. Kentner, D.; Martano, G.; Callon, M.; Chiquet, P.; Brodmann, M.; Burton, O.; Wahlander, A.; Nanni, P.; Delmotte, N.; Grossmann, J.; et al. Shigella reroutes host cell central metabolism to obtain high-flux nutrient supply for vigorous intracellular growth. Proc. Natl. Acad. Sci. USA 2014, 111, 9929–9934. [Google Scholar] [CrossRef] [Green Version]
  88. Pandey, A.K.; Sassetti, C.M. Mycobacterial persistence requires the utilization of host cholesterol. Proc. Natl. Acad. Sci. USA 2008, 105, 4376–4380. [Google Scholar] [CrossRef] [Green Version]
  89. Gierok, P.; Harms, M.; Methling, K.; Hochgräfe, F.; Lalk, M. Staphylococcus aureus infection reduces nutrition uptake and nucleotide biosynthesis in a human airway epithelial cell line. Metabolites 2016, 6, 41. [Google Scholar] [CrossRef]
  90. Bravo-Santano, N.; Ellis, J.K.; Calle, Y.; Keun, H.C.; Behrends, V.; Letek, M. Intracellular Staphylococcus aureus Elicits the Production of Host Very Long-Chain Saturated Fatty Acids with Antimicrobial Activity. Metabolites 2019, 9, 148. [Google Scholar] [CrossRef] [Green Version]
  91. Zhu, Y.; Li, H.; Ding, S.; Wang, Y. Autophagy inhibition promotes phagocytosis of macrophage and protects mice from methicillin-resistant Staphylococcus aureus pneumonia. J. Cell. Biochem. 2018, 119, 4808–4814. [Google Scholar] [CrossRef]
  92. Rossi, J.-F.; Lu, Z.-Y.; Jourdan, M.; Klein, B. Interleukin-6 as a Therapeutic Target. Clin. Cancer Res. 2015, 21, 1248–1257. [Google Scholar] [CrossRef] [PubMed]
  93. Rahman, S.; Rehn, A.; Rahman, J.; Andersson, J.; Svensson, M.; Brighenti, S. Pulmonary tuberculosis patients with a vitamin D deficiency demonstrate low local expression of the antimicrobial peptide LL-37 but enhanced FoxP3+ regulatory T cells and IgG-secreting cells. Clin. Immunol. 2015, 156, 85–97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Napier, R.J.; Norris, B.A.; Swimm, A.; Giver, C.R.; Harris, W.A.C.; Laval, J.; Napier, B.A.; Patel, G.; Crump, R.; Peng, Z.; et al. Low Doses of Imatinib Induce Myelopoiesis and Enhance Host Anti-microbial Immunity. PLoS Pathog. 2015, 11, e1004770. [Google Scholar] [CrossRef] [PubMed]
  95. Sinha, B.; François, P.P.; Nüsse, O.; Foti, M.; Hartford, O.M.; Vaudaux, P.; Foster, T.J.; Lew, D.P.; Herrmann, M.; Krause, K.H. Fibronectin-binding protein acts as Staphylococcus aureus invasin via fibronectin bridging to integrin alpha5beta1. Cell. Microbiol. 1999, 1, 101–117. [Google Scholar] [CrossRef]
  96. Agerer, F.; Lux, S.; Michel, A.; Rohde, M.; Ohlsen, K.; Hauck, C.R. Cellular invasion by Staphylococcus aureus reveals a functional link between focal adhesion kinase and cortactin in integrin-mediated internalisation. J. Cell Sci. 2005, 118, 2189–2200. [Google Scholar] [CrossRef] [Green Version]
  97. Richter, E.; Harms, M.; Ventz, K.; Nölker, R.; Fraunholz, M.J.; Mostertz, J.; Hochgräfe, F. Quantitative Proteomics Reveals the Dynamics of Protein Phosphorylation in Human Bronchial Epithelial Cells during Internalization, Phagosomal Escape, and Intracellular Replication of Staphylococcus aureus. J. Proteome Res. 2016, 15, 4369–4386. [Google Scholar] [CrossRef]
  98. Goldmann, O.; Tuchscherr, L.; Rohde, M.; Medina, E. α-Hemolysin enhances Staphylococcus aureus internalization and survival within mast cells by modulating the expression of β1 integrin. Cell. Microbiol. 2015, 18, 807–819. [Google Scholar] [CrossRef]
  99. Ashraf, S.; Cheng, J.; Zhao, X. Clumping factor A of Staphylococcus aureus interacts with AnnexinA2 on mammary epithelial cells. Sci. Rep. 2017, 7, 40608. [Google Scholar] [CrossRef] [Green Version]
  100. Oviedo-Boyso, J.; Cortés-Vieyra, R.; Huante-Mendoza, A.; Yu, H.B.; Valdez-Alarcón, J.J.; Bravo-Patiño, A.; Cajero-Juárez, M.; Finlay, B.B.; Baizabal-Aguirre, V.M. The phosphoinositide-3-kinase-akt signaling pathway is important for Staphylococcus aureus internalization by endothelial cells. Infect. Immun. 2011, 79, 4569–4577. [Google Scholar] [CrossRef] [Green Version]
  101. McDonnell, C.J.; Garciarena, C.D.; Watkin, R.L.; McHale, T.M.; McLoughlin, A.; Claes, J.; Verhamme, P.; Cummins, P.M.; Kerrigan, S.W. Inhibition of major integrin αVβ3 reduces Staphylococcus aureus attachment to sheared human endothelial cells. J. Thromb. Haemost. 2016, 14, 2536–2547. [Google Scholar] [CrossRef]
  102. Ellington, J.K.; Elhofy, A.; Bost, K.L.; Hudson, M.C. Involvement of Mitogen-Activated Protein Kinase Pathways in Staphylococcus aureus Invasion of Normal Osteoblasts. Infect. Immun. 2001, 69, 5235–5242. [Google Scholar] [CrossRef] [Green Version]
  103. Soong, G.; Martin, F.J.; Chun, J.; Cohen, T.S.; Ahn, D.S.; Prince, A. Staphylococcus aureus protein A mediates invasion across airway epithelial cells through activation of RhoA GTPase signaling and proteolytic activity. J. Biol. Chem. 2011, 286, 35891–35898. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Zhao, S.; Gao, Y.; Xia, X.; Che, Y.; Wang, Y.; Liu, H.; Sun, Y.; Ren, W.; Han, W.; Yang, J.; et al. TGF-β1 promotes Staphylococcus aureus adhesion to and invasion into bovine mammary fibroblasts via the ERK pathway. Microb. Pathog. 2017, 106, 25–29. [Google Scholar] [CrossRef] [PubMed]
  105. Dziewanowska, K.; Carson, A.R.; Patti, J.M.; Deobald, C.F.; Bayles, K.W.; Bohach, G.A. Staphylococcal fibronectin binding protein interacts with heat shock protein 60 and integrins: Role in internalization by epithelial cells. Infect. Immun. 2000, 68, 6321–6328. [Google Scholar] [CrossRef] [PubMed]
  106. Hirschhausen, N.; Schlesier, T.; Schmidt, M.A.; Götz, F.; Peters, G.; Heilmann, C. A novel staphylococcal internalization mechanism involves the major autolysin Atl and heat shock cognate protein Hsc70 as host cell receptor. Cell. Microbiol. 2010, 12, 1746–1764. [Google Scholar] [CrossRef] [PubMed]
  107. Askarian, F.; Ajayi, C.; Hanssen, A.-M.; van Sorge, N.M.; Pettersen, I.; Diep, D.B.; Sollid, J.U.E.; Johannessen, M. The interaction between Staphylococcus aureus SdrD and desmoglein 1 is important for adhesion to host cells. Sci. Rep. 2016, 6, 22134. [Google Scholar] [CrossRef] [Green Version]
  108. Yang, Y.-H.; Jiang, Y.-L.; Zhang, J.; Wang, L.; Bai, X.-H.; Zhang, S.-J.; Ren, Y.-M.; Li, N.; Zhang, Y.-H.; Zhang, Z.; et al. Structural Insights into SraP-Mediated Staphylococcus aureus Adhesion to Host Cells. PLoS Pathog. 2014, 10, e1004169. [Google Scholar] [CrossRef]
  109. Miller, M.; Dreisbach, A.; Otto, A.; Becher, D.; Bernhardt, J.; Hecker, M.; Peppelenbosch, M.P.; van Dijl, J.M. Mapping of Interactions between Human Macrophages and Staphylococcus aureus Reveals an Involvement of MAP Kinase Signaling in the Host Defense. J. Proteome Res. 2011, 10, 4018–4032. [Google Scholar] [CrossRef]
  110. Inoshima, I.; Inoshima, N.; Wilke, G.A.; Powers, M.E.; Frank, K.M.; Wang, Y.; Wardenburg, J.B. A Staphylococcus aureus pore-forming toxin subverts the activity of ADAM10 to cause lethal infection in mice. Nat. Med. 2011, 17, 1310–1314. [Google Scholar] [CrossRef] [Green Version]
  111. Neumann, Y.; Bruns, S.A.; Rohde, M.; Prajsnar, T.K.; Foster, S.J.; Schmitz, I. Intracellular Staphylococcus aureus eludes selective autophagy by activating a host cell kinase. Autophagy 2016, 12, 2069–2084. [Google Scholar] [CrossRef] [Green Version]
  112. Schröder, A.; Schröder, B.; Roppenser, B.; Linder, S.; Sinha, B.; Fässler, R.; Aepfelbacher, M. Staphylococcus aureus Fibronectin Binding Protein-A Induces Motile Attachment Sites and Complex Actin Remodeling in Living Endothelial Cells. Mol. Biol. Cell 2006, 17, 5198–5210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Boada-Romero, E.; Letek, M.; Fleischer, A.; Pallauf, K.; Ramó n-Barros, C.; Pimentel-Muiñ os, F.X. TMEM59 defines a novel ATG16L1-binding motif that promotes local activation of LC3. EMBO J. 2013, 32, 566–582. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Mestre, M.B.; Colombo, M.I. CAMP and EPAC are key players in the regulation of the signal transduction pathway involved in the α-hemolysin autophagic response. PLoS Pathog. 2012, 8, e1002664. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Imre, G.; Heering, J.; Takeda, A.-N.; Husmann, M.; Thiede, B.; zu Heringdorf, D.M.; Green, D.R.; van der Goot, F.G.; Sinha, B.; Dötsch, V.; et al. Caspase-2 is an initiator caspase responsible for pore-forming toxin-mediated apoptosis. EMBO J. 2012, 31, 2615–2628. [Google Scholar] [CrossRef] [PubMed]
  116. Rudel, T.; Kepp, O.; Kozjak-Pavlovic, V. Interactions between bacterial pathogens and mitochondrial cell death pathways. Nat. Rev. Microbiol. 2010, 8, 693. [Google Scholar] [CrossRef] [PubMed]
  117. Muñoz-Planillo, R.; Franchi, L.; Miller, L.S.; Núñez, G. A critical role for hemolysins and bacterial lipoproteins in Staphylococcus aureus-induced activation of the Nlrp3 inflammasome. J. Immunol. 2009, 183, 3942–3948. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Table 1. List of host-directed drugs able to halt the intracellular infection caused by different pathogens.
Table 1. List of host-directed drugs able to halt the intracellular infection caused by different pathogens.
ClassDrugMechanism of ActionPathogenReference
Monoclonal antibodyAdalimumabAnti-TNFαMycobacterium tuberculosis[38]
Anti-interleukin 1βCytokine neutralisationHelicobacter pylori[39]
Antipertussis toxins antibodyEnhancement of immunoglobulinsBordetella pertussis[40]
Anti-TNFαCytokine neutralisationHelicobacter pylori[39]
BevacizumabAnti-VEGFMycobacterium tuberculosis[41,42]
NivolumabAnti-PD-1Mycobacterium tuberculosis[43,44]
SiltuximabAnti-interleukin 6Mycobacterium tuberculosis[45]
Repurposed drugAspirinNSAID, TNFα levels reductionMycobacterium tuberculosis[46]
ChlorpromazineCalmodulin antagonistStaphylococcus aureus[47]
Neorickettsia risticii[48]
Salmonella Typhimurium[49]
ETB067Serine-threonine protein kinase (Akt1) inhibitorMycobacterium tuberculosis[50]
FingolimodActivation of sphingosine-1-phosphate pathwayBordetella pertussis[51]
GlibendamideCyclooxygenase inhibitionStreptococcus pneumoniae[52]
H-89Protein kinase A (PKA) inhibitorSalmonella Typhimurium[50]
Coxiella burnetii[53]
IbuprofenNSAID, cyclooxygenase inhibitionStreptococcus pneumoniae[54]
Mycobacterium tuberculosis[55,56]
Imatinib mesylateBCR-ABL tyrosine kinase inhibitorMycobacterium tuberculosis[57]
Anaplasma phagocytophilum[58]
Repurposed drugIndometacinCyclooxygenase inhibitionStreptococcus pneumoniae[59]
MetforminMitochondrial respiratory chain blockerMycobacterium tuberculosis[60]
NiraparibPARP inhibitorMycobacterium tuberculosis[61]
PhenylbutyrateHistone deacetylase inhibitorMycobacterium tuberculosis[62]
PimozideCalcium channel inhibitorListeria monocytogenes[63]
Bacillus subtilis[63]
Salmonella Typhimurium[63]
Escherichia coli[63]
PrednisoneGlucocorticoid receptor antagonistStreptococcus pneumoniae[64]
Mycobacterium tuberculosis[65]
RaloxifeneOestrogen receptor modulatorPseudomonas aeruginosa[66]
StatinsHMG-CoA reductase inhibitorStreptococcus pneumoniae[67,68]
Mycobacterium tuberculosis[69]
SulforaphaneHistone deacetylase inhibitorNeisseria gonorrhoeae[70]
ThapsigarginCalcium ATPase inhibitorCoxiella burnetii[53]
ThioridazineunknownListeria monocytogenes[71]
Staphylococcus aureus[47]
Mycobacterium tuberculosis[72]
VerapamilCalcium channel inhibitorMycobacterium tuberculosis[73,74]
Chlamydia pneumoniae[75]
VorinostatHistone deacetylase inhibitorMycobacterium tuberculosis[76]
ZileutonLeukotriene synthesis inhibitorMycobacterium tuberculosis[77]
VitaminVitamin D3Activation of antimicrobial defensesHelicobacter pylori[78]
Table 2. Host molecular factors hijacked by S. aureus during intracellular infection and potential host-directed drugs against intracellular S. aureus.
Table 2. Host molecular factors hijacked by S. aureus during intracellular infection and potential host-directed drugs against intracellular S. aureus.
Host FactorPutative FunctionReferenceHost-Directed DrugType
Adherence and Internalization
α5β1-integrins Internalization into non-phagocytic cells[95]Volociximab *Antibody
FAKInternalization into non-phagocytic cells[96,97]PF-562271Inhibitor
Src-mediated cortactinInternalization into non-phagocytic cells[96,97]PP2Inhibitor
β1-integrins Internalization into mast cells[98] R1295 *Antagonist
Annexin 2Internalization into epithelial cells[99]
(PI3K)-Akt Internalization into endothelial cells[100] Nelfinavir *Inhibitor
αVβ3-integrin Internalization into endothelial cells[101] Cilengitide *Inhibitor
ERKInternalization into osteoblast and Hep-2 cells[102] SCH772984 *Inhibitor
ERK1/2/MEKPenetration into airway epithelial cells[103] UO126Inhibitor
ERKInvasion to fibroblasts[104] PD98059Inhibitor
Hsp60Internalization into epithelial cells[105]
Hsc70Internalization into 293T cells[106]
Desmoglein 1 Adherence to keratinocytes[107]
Scavenger protein gp340 Internalization into A549 cells [108]
EGFRPenetration into airway epithelial cells[103]BPDQInhibitor
ROCKPenetration into airway epithelial cells[103]Y-27632Inhibitor
JNKPenetration into airway epithelial cells[103]SP600125Inhibitor
p38/MAPKPenetration into airway epithelial cells[103]SB202190Inhibitor
EPHA2Invasion/Internalization into epithelial cells[34]IbrutinibInhibitor
CDKAdhesion to human bronchial epithelial cells[97]Roscovitine
PKAInternalization into Thp1 macrophages[97,109]H-89Inhibitor
PKCInternalization into Thp1 macrophages[97,109]Bisindolylmaleimide-IInhibitor
Intracellular Survival and Proliferation
ADAM10Cleavage of adherens junction protein E-cadherin[110]GI 254023X Inhibitor
AMPKInduction of autophagy[17]DorsomorphinInhibitor
ERKInduction of autophagy[17]SCH772984 *Inhibitor
TRAM2Ca2+ pump to promote collagen synthesis[81]ThapsigarginInhibitor
p38/MAPKSubversion of autophagy[111]Skepinone-L *Inhibitor
PAKCytoeskeleton rearrangements[97]FRAX597 *Inhibitor
MYL2Cytoeskeleton rearrangements[81]Blebbistatin *Inhibitor
FAM63BIntracellular trafficking[81]
ActinPromote bacterial movements within the host cell[95,112]Cytochalasin D Inhibitor
Rab5Promote bacterial movements within the host cell[112]
NWASPProduction of actin-comet tails to facilitate movement[112]WiskostatinInhibitor
TMEM59Activation of selective-autophagy[113]
RAPGEF3Induction of autophagy[114]Salirasib *Inhibitor
RAP2BInduction of autophagy[114]
S. aureus-Induced Host Cell Death
Caspase 2S. aureus-induced apoptosis[115]Z-VDVAD-FMKInhibitor
Caspase 9S. aureus-induced apoptosis[116]Z-LEHD-FMKInhibitor
NLRP3S. aureus-induced pyronecrosis[117]MCC950 *Inhibitor
*: Drugs that could potentially halt intracellular S. aureus by targeting the described host pathway, but their effects have not been yet investigated.

Share and Cite

MDPI and ACS Style

Bravo-Santano, N.; Behrends, V.; Letek, M. Host-Targeted Therapeutics against Multidrug Resistant Intracellular Staphylococcus aureus. Antibiotics 2019, 8, 241. https://0-doi-org.brum.beds.ac.uk/10.3390/antibiotics8040241

AMA Style

Bravo-Santano N, Behrends V, Letek M. Host-Targeted Therapeutics against Multidrug Resistant Intracellular Staphylococcus aureus. Antibiotics. 2019; 8(4):241. https://0-doi-org.brum.beds.ac.uk/10.3390/antibiotics8040241

Chicago/Turabian Style

Bravo-Santano, Natalia, Volker Behrends, and Michal Letek. 2019. "Host-Targeted Therapeutics against Multidrug Resistant Intracellular Staphylococcus aureus" Antibiotics 8, no. 4: 241. https://0-doi-org.brum.beds.ac.uk/10.3390/antibiotics8040241

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