4,522
Views
9
CrossRef citations to date
0
Altmetric
Review

The role of the human gut microbiota in colonization and infection with multidrug-resistant bacteria

, , &
Article: 1911279 | Received 29 Jun 2020, Accepted 23 Mar 2021, Published online: 18 Apr 2021

ABSTRACT

About 100 years ago, the first antibiotic drug was introduced into health care. Since then, antibiotics have made an outstanding impact on human medicine. However, our society increasingly suffers from collateral damage exerted by these highly effective drugs. The rise of resistant pathogen strains, combined with a reduction of microbiota diversity upon antibiotic treatment, has become a significant obstacle in the fight against invasive infections worldwide.

Alternative and complementary strategies to classical “Fleming antibiotics” comprise microbiota-based treatments such as fecal microbiota transfer and administration of probiotics, live-biotherapeutics, prebiotics, and postbiotics. Other promising interventions, whose efficacy may also be influenced by the human microbiota, are phages and vaccines. They will facilitate antimicrobial stewardship, to date the only globally applied antibiotic resistance mitigation strategy.

In this review, we present the available evidence on these nontraditional interventions, highlight their interaction with the human microbiota, and discuss their clinical applicability.

Introduction

The human body harbors a multitude of microorganisms, including bacteria, fungi, archaea, and viruses, which exist in a symbiotic relationship with their host. The entirety of these commensals is referred to as the microbiota, and their collective genomic information as the human microbiome.Citation1,Citation2 Within the microbiota, bacteria play a central role. All body surfaces are characterized by their specific bacteriome, which describes the bacterial component of the microbiome. Over 2000 bacterial species have been identified as human commensals, a majority of which remain uncultured.Citation3 The gut microbiota composition varies between individual persons and has been found to consist of a few hundred bacterial operational taxonomic units (OTU) on average.Citation4–6 They constitute a subset of the overall phylogenetic diversity found in the corresponding human population.Citation7–9 Most of these bacteria reside in the colonCitation10 and occupy different functional niches.Citation3

The development of high-throughput sequencing techniques has improved our understanding of the role our commensal bacteria play in maintaining human homeostasis. Their regulatory properties are central to many physiological processes associated with health and disease, and highly diverse in nature.Citation11–13 Several functional axes have been identified over the last years, e.g., the gut–brain axis, the gut–liver axis, the gut–lung axis, and the gut–immune axis.Citation14–17 Moreover, direct and indirect effects of the human microbiome on bacterial infection are varied and complexCitation18,Citation19 (). The human microbiota is increasingly recognized as a therapeutic target for infection prevention and treatment.

Figure 1. Intervention strategies against multidrug-resistant bacterial pathogens that are mediated or boosted by healthy commensal microbiota. In cases of dysbiosis or dysregulation, the microbiota may also contribute to increased pathogen colonization and disease. FMT: fecal microbiota transfer

Figure 1. Intervention strategies against multidrug-resistant bacterial pathogens that are mediated or boosted by healthy commensal microbiota. In cases of dysbiosis or dysregulation, the microbiota may also contribute to increased pathogen colonization and disease. FMT: fecal microbiota transfer

Direct effects of microbial commensals: colonization resistance and pathobionts

Colonization resistance refers to the protection by the healthy microbiota against host colonization with pathogenic microorganisms. The search for the definition of a healthy microbiota has not been concluded, but a multitude of independent findings confirm that a high alpha diversity is associated with good health.Citation20,Citation21 The microbiota may however be a reservoir for potentially pathogenic commensals that can turn into causative agents of endogenic infections, so-called pathobionts. A loss in diversity or a disproportionate increase in one or more commensal species often indicates the presence of a disease state. Such shifts are commonly referred to as dysbiosis, even though an exact definition of this term remains to be established. The absence of dysbiosis plays a crucial role in the functionality of colonization resistance.

Individuals become colonized with multidrug-resistant (MDR) bacteria through contact with the healthcare system, the environment, animals, or the food chain. Initially, these bacteria may be present below the level of detection. However, exposure to antibiotics or other substances that exert selection pressure on the microbiota facilitates rapid expansion and domination of MDR bacteria. If these changes coincide with a breach in host barrier functions, e.g., in the context of chemotherapy or surgery, bacterial translocation, and infection with MDR bacteria become highly likely.

Different studies support this model of infection pathogenesis. Mice gavaged with vancomycin-resistant enterococci (VRE) prior to antibiotic exposure displayed functional colonization resistance with only minimal amounts of VRE detectable in their gut microbiota. If gavaged after antibiotic exposure, however, VRE was able to successfully colonize the gut even weeks after exposure. Apparently, some antibiotics are able to open a niche in the gut that favors the survival of VRE.Citation22 With respect to Gram-negative MDR bacteria, an understanding of the niche required for the growth of MDR bacteria is less well established. Recent findings suggest that the synthesis of short-chain fatty acids (SCFA) by gut commensals may play an important role in this setting. While a balanced gut microbiota synthesizes enough SCFA to maintain an acidic pH in the gut, exposure to antibiotics induces a dysbiosis in the gut microbiota that leads to decreased SCFA production and an increase in pH. Under these circumstances, Gram-negative MDR bacteria are more likely to colonize and dominate the gut.Citation23

Indirect effects of microbial commensals: immune system regulation

The microbiome profoundly influences the host’s immune system. At birth, the innate and adaptive immune system is not yet fully developed. Interactions with microbes provide a central role in their development process by direct contact with commensal symbionts in addition to environmental antigens.Citation24 During the first years of life of an infant, the early microbiota can shape the immune system and vice versa.Citation25 The interactions between commensal bacteria and the human immune system are complex and their study is still in its infancy.

On the epithelial surface, the microbiota can regulate the integrity of the epithelial barrier, thereby preventing the penetration of pathogens into the host tissue and blood stream, and help respond to epithelial damage and pathogen breaches.Citation26,Citation27 This is mediated via pattern-recognition receptors (PRR) displayed in epithelial, endothelial, and immune cells, which detect microbe-associated molecular patterns (MAMPs) such as lipopolysaccharides and flagellin.Citation28,Citation29 Subsequent production of chemokines recruits immune cells, or activates the inflammasome.Citation27,Citation30,Citation31 In turn, aspects of the inflammasome may influence the commensal microbiota.Citation32 MAMPs are, however, not unique to pathogens, and it is unclear how commensals are distinguished from pathogens.Citation26 Location may play an important role: commensals are mostly sequestrated on epithelial surfaces, while pathogens cross the epithelial barrier.Citation33,Citation34

Metabolites produced by bacterial commensals modulate innate and adaptive immune cells.Citation35 A metabolite may have different effects depending on the receptor cell type, such as differentiation, activation, inhibition, migration, or production of antimicrobial peptides.Citation36–39

Immune tolerance is the state of unresponsiveness of the immune system to agents that have the potential to induce an immune response.Citation40,Citation41 Intestinal regulatory T (Treg) cells play pivotal roles in the suppression of immune responses against harmless dietary antigens and commensal microorganisms.Citation42 Differentiation, localization, and maintenance of intestinal Treg cells and tolerogenic dendritic cells are controlled by signals from the intestinal microbiota.Citation43 Molecular mechanisms may involve PRR signaling or generation of microbial metabolites, but are still largely unidentified.

Finally, the commensal microbiota can contribute to dysregulation of the immune response. Bacterial dysbiosis has been associated with various diseases, among them asthma, allergies, obesity, chronic inflammatory disorders of the skin, colorectal cancer, and cardiovascular disease.Citation44–47 In addition, dysbiosis of the commensal microbiota, such as caused by antibiotic treatment, has been associated with an increased risk of bloodstream infection and sepsis.Citation19,Citation48

Alternative interventions against MDR pathogens

Microbiota-based treatments

Based on our understanding of the key role of the microbiota in the prevention of colonization and infection with MDR bacteria, different microbiota-based treatments can be envisioned. Efforts to engineer or influence the commensal microbiota can be divided into two strategies: administration of live microorganisms, and supplementation with factors that influence the commensal microbiota.

Fecal microbiota transplantation

Fecal microbiota transplantation (FMT) is the transfer of stool from a healthy donor to the intestine of a patient. This can either be done endoscopically, by rectal enema, by oral ingestion of encapsulated preparations or by nasogastric or nasoduodenal tube. FMT is currently only recommended for the treatment of recurrent Clostridioides difficile infection (CDI), with an efficacy rate of up to 90%.Citation49,Citation50 How FMT decolonizes C. difficile has not been fully established, and microbial predictors of therapeutic outcome are not clear. Bile acid metabolism seems to play one of the central roles, however; the primary bile acid taurocholate, secreted by the liver, induces germination of C. difficile spores.Citation51 Certain members of the commensal microbiota are able to metabolize primary bile acids and convert them into secondary bile acids, such as deoxycholate. While also able to induce germination, deoxycholate inhibits the vegetative growth of C. difficile.Citation51,Citation52 Other potential mechanisms for FMT efficacy may include microbiota modulation by direct interaction or competition (including quorum sensing modulation), and host immunity modulation.Citation53

FMT is not currently approved for clinical use in the USA, but considered an investigational new drug.Citation54 It is regulated individually in other countries.Citation55 FMT is subject to safety concerns, namely transmission of infectious agents, including MDR pathogens,Citation56 and unidentified risks associated with changes in the patient’s microbiota. Improved regulation, manufacturing standards, and stool banks are expected to mitigate the former.

In patients who receive FMT, a significant reduction in fecal bacterial antibiotic resistance genes was observed for Gram-positive pathogens,Citation57,Citation58 which suggests that FMT may harbor the potential to displace multiresistant bacteria from intestinal microbiota.Citation59 Some clinical data are already available in this respect: in a recent study investigating the efficacy of a microbiota preparation as a treatment for recurrent CDI, successful VRE decolonization was recorded as a side effect.Citation60 Similar results were documented in another small case study.Citation61

With regard to decolonization of Gram-negative MDR bacteria, several case reports and uncontrolled studies, as well as one randomized trial assessing this question, provide mixed results to date.Citation62 Application of FMT for MDR pathogen suppression in solid organ transplant recipients and patients with hematologic malignancies resulted in partial or full decolonization in multiple cases.Citation63 Other recent anecdotal cases report eradication of MDR Klebsiella pneumoniae in a critically ill patient with endocarditis and sepsis originating from an infection of a pacemaker component,Citation64 and MDR pathogen elimination in the case of cholangitis (inflammation of the bile duct system) with associated bacteremia.Citation65

FMT is explored for treatment of other intestine-associated complications, such as inflammatory bowel disease, ulcerative colitis, and Crohn’s disease.Citation66 Moreover, insights into the manifold effects of microbiota on human metabolism have spurred experimental FMT therapy for indications such as bipolar disorder (NCT03279224), Parkinson's disease (NCT03808389), cirrhosis (NCT02862249), and psoriatic arthritis (NCT03058900). Randomized controlled trials are currently ongoing, as well as experimental treatment of other diseases such as autoimmune diseases.Citation67

Probiotics and live biotherapeutics

Probiotics are viable microorganisms which, when administered in sufficient quantities, have beneficial effects on the health of the host.Citation68 If used as a drug with an associated health claim, they are referred to as live biotherapeutic products/agents.Citation69 While probiotics are traditionally isolated from food, live biotherapeutics may be isolated from various niches. The latter may also include genetically modified organisms. Disease targets range from cancer, to autoimmune diseases (including asthma), to clearance of infectious agents. Mechanisms of action are specific to individual strains, and generally fall into one or multiple categories: microbiota modulation by direct interaction or competition, host metabolism modification,Citation11,Citation12 and host immunity modulation.

An example of a direct colonization resistance mechanism comes from Staphylococcus lugdunensis, a commensal isolated from the human nose.Citation70 It produces lugdunin, a cyclic peptide antibiotic that inhibits the growth of various Gram-positive pathogens, including VRE and methicillin-resistant Staphylococcus aureus. Moreover, lugdunin amplifies the innate immune response by inducing expression of antimicrobial peptides and pro-inflammatory chemokines in human keratinocytes.Citation71

Efforts are currently being made to identify defined bacterial consortia to suppress MDR strains.Citation18,Citation72 Importantly, administered strains must be free of antibiotic resistance genes.Citation73 Results of randomized trials for MDR decolonization are mixed to date: two studies using Lactobacillus rhamnosus GG showed success in decolonizing patients with VRE.Citation74,Citation75 No effect on colonization was achieved against various Gram-negative MDR pathogens using a combination of Lactobacillus bulgaricus and L. rhamnosus.Citation76

Prebiotics and postbiotics

Prebiotics are non-digestible food components that favorably influence human health by modulating the growth and/or activity of one or more species of commensals. Metabolites and cell components derived from probiotic strains, which influence the microbiota and host health, are referred to as postbiotics.Citation77

Among the most commonly used prebiotics are oligosaccharides such as inulin, fructo-oligosaccharides and galacto-oligosaccharides: their fermentation by gut microbiota results in SCFA. Other classes of prebiotics are human milk oligosaccharides (HMO), conjugated linoleic acid and polyunsaturated fatty acids, polyphenols, and fermentable dietary fibers.Citation78,Citation79 The health benefits of prebiotics mostly depend on microbial utilization and the metabolites produced, rather than on parent compounds.Citation80

Prebiotics may help prevent dysbiosis. In healthy volunteers, who were exposed to antibiotics, synthetically produced HMO positively influenced restoration of a balanced microbiota by selectively stimulating the growth of certain species, e.g., Actinobacteria and Bacteroidetes, while suppressing others, e.g., Firmicutes.Citation81 HMO can also influence the innate immune response, and directly prevent adhesion of pathogens to the intestinal epithelium.Citation82,Citation83

Postbiotic metabolites comprise enzymes, proteins and peptides, lipids, polysaccharides, and organic acids. Postbiotic components include peptidoglycan, cell-surface proteins, exopolysaccharides, and teichoic acids. Postbiotics can act directly on the host (e.g. immunomodulation), the microbiota, or colonizing pathogens. In mice, extracellular vesicles from Akkermansia muciniphila and probiotic Escherichia coli increased the integrity of the epithelial gut barrier, contributing to one of the crucial factors that prevents systemic infection.Citation84,Citation85

Phage-mediated therapies

Phages are viral bodies that infect bacteria via attachment to bacterial surface-proteins and introduce their own DNA into the bacterial genome. They exploit the bacterial transcription and translation machinery for the production of infectious particles. After assembly, phages exit via lysis,Citation86 leading to cycles of reinfection and phage-mediated genome exchange between bacterial hosts. Microbiomes and phages are directly dependent on each other and in a state of continuous co-evolution.Citation87 Coinfections of bacteria with multiple phages are the norm, resulting in a dynamic network of horizontal gene transfer that includes antibiotic resistance genes. While these gene transfers may result in the spread of resistance genes within a host, the mechanisms and the clinical relevance of these dynamics are topics of controversial discussions.Citation88–90

The principle of bacterial genome-modification and lysis also holds therapeutic promise. D’Herelle et al. showed first-in-human application in Vibrio cholerae infection in the early twentieth century.Citation91,Citation92 Since then, the characterization of phages and phage–host interactions have been studied in depth to enable the translation of these findings into clinical practice.Citation93–95 A key feature of phages is their high host specificity.Citation96 Most known phages infect only a few strains of closely related bacterial populations, which leaves most of the commensal bacteria undisturbed. A variety of animal studies were able to show that phages can be used to eliminate MDR bacteria, including MDR P. aeruginosa, A. baumannii, and VRE.Citation97–99

A number of smaller randomized controlled trials assessing the efficacy of phages in the treatment of bacterial infections have been published. While none of these studies reported any problems with respect to safety, response to treatment was inconsistent between studies.Citation100–105 Next to further improvement of efficacy, a regulatory framework for phage therapies needs to be put in place.

Vaccines and antibodies

Vaccines and antibodies are designed to prevent infection or to decrease disease severity. Effective vaccines deliver antigens to elicit a prophylactic immune response to generate disease-specific antibodies with their corresponding memory cells, and provide long-term protection against invading pathogens.Citation106,Citation107 Externally administered antibodies are agents of passive immunization, act faster than vaccines (hours or days), and bestow short-term protection.Citation108

Neither vaccines nor antibody preparations specifically act against MDR bacterial strains to date, but they can reduce the spread of targeted pathogenic strains.Citation109 This may decrease the number of antibiotics used, which reduces the selection pressure on antibiotic resistance, aligns with antimicrobial stewardship measures, and forgoes antibiotic de-colonization of commensal microbiota. By avoiding antibiotic treatment, vaccines can potentially reduce bystander selection of resistance elements in the commensal microbiota: the proliferation of bacteria carrying resistance genes upon targeted antibiotic removal of a neighboring pathogenic species.Citation110,Citation111

Resistance development to vaccines is relatively infrequent compared to the emergence of antibiotic resistance,Citation112 but serotype replacement followed by the spread of new MDR serotypes have been observed,Citation113 as well as an increase in invasive, non-vaccine serotype strains of targeted pathogens (e.g. Haemophilus influenzae, Streptococcus pneumoniae).Citation114–116 Pathogenic strains are often heterogeneous and diverse, which increases the complexity of vaccine development.Citation117 Moreover, antibiotic resistance mechanisms are frequently encoded in mobile genetic elements and horizontally transferred.Citation118,Citation119 Vaccines directed at the effectors of drug resistance, such as penicillin-binding proteins and β-lactamases, are being studied in animal models.Citation120,Citation121 Vaccines against gut-associated bacterial pathogens are currently available for Vibrio cholerae (Vaxchora), Salmonella typhi (Vivotif Berna), as well as Bacillus anthracis (BioThrax).

The commensal microbiota was shown to influence vaccine efficacy. Flagellin derived from commensal microbiota, for instance, may play an adjuvant role and enhance immune response in response to vaccination.Citation122,Citation123 Components of lipoteichoic acid and peptidoglycan appear to have similar effects.Citation124,Citation125

Mitigating antibiotic collateral damage

Approximately, a quarter of all inpatients treated at hospitals receive antibiotics. One-third to one-half of antibiotic prescriptions in inpatient settings are insufficient with regard to the indication and/or duration of therapy.Citation126,Citation127 This fosters selection for antibiotic resistance and the spread of nosocomial MDR strains. The use of broad-spectrum antibiotics can devastate the beneficial commensal microbiota, thus rendering a patient temporarily more susceptible to opportunistic infections, induce dysbiosis, and even cause long-lasting complications such as asthma and inflammatory bowel disease.Citation128–130 Potential solutions to address this issue are the development of treatment options with increased specificity or confined activity. Examples are narrow-spectrum antibiotics,Citation131 selective pathogen-targeting phagesCitation96 or antibody-antibiotic conjugates,Citation132 localized antibiotic delivery using nanoparticle formulations,Citation133 and strategies to protect the gut microbiota with orally administered beta-lactamasesCitation134 or slow-release formulations of activated charcoal that absorb antibiotics as soon as they reach the large intestine.Citation134

Antimicrobial stewardship

Antimicrobial stewardship (AMS) programs constitute a vital strategy to address antimicrobial resistance. They are designed to improve the quality of antimicrobial prescriptions in terms of substance selection, dosage, route of administration, and duration of therapy. AMS has been shown to reduce adverse events such as sepsis, results in lower mortality rates and improves patient outcomes, and decreases the rate of colonization and infection with MDR bacteria.Citation135–138 The above outlined potential alternative therapies are not only fully compatible with AMS strategies, but would facilitate its implementation and potentially enhance its effectiveness.

Conclusion and outlook

As the antibiotic resistance crisis is unfolding, it has become clear that we cannot rely on classic antibiotics alone to suppress the rise of MDR bacterial pathogens. While antibiotic stewardship and the development of single-component molecule antibiotics are of utmost importance, we will increasingly have to rely on alternative strategies to support or even replace conventional antibiotic treatment. Vaccines and phages have already shown great promise in past applications and proof-of-concept studies. Our greatest ally may yet become the human commensal microbiota. While it is only recently that we began to systematically uncover the myriad ways in which these microorganisms contribute to health and disease, it is already evident that they offer many intervention points to combat infectious agents directly or in tandem with the human host. We have a lot of work ahead of us to attain safe and efficacious treatments, and it is most certainly exciting and promising.

Disclosure of interest

MJGTV has received research grants from 3M, Astellas Pharma, DaVolterra, Gilead Sciences, Glycom, MaaT Pharma, Merck/MSD, Organobalance, Seres Therapeutics; speaker fees from Astellas Pharma, Basilea, Gilead Sciences, Merck/MSD, Organobalance, Pfizer and has been a consultant to Alb Fils Kliniken GmbH, Astellas Pharma, DaVolterra, Ferring, MaaT Pharma, Merck/MSD. YK has received speaker fees from Merck/MSD. IW and BP report no conflict of interest.

Author contributions

IW, BP, YK and MJGTV wrote sections of the manuscript. All authors contributed to manuscript revision, read and approved the submitted version.

Additional information

Funding

No funding to declare.

References

  • Arumugam M, Raes J, Pelletier E, Le Paslier D, Yamada T, Mende DR, Fernandes GR, Tap J, Bruls T, Batto JM, et al. Enterotypes of the human gut microbiome. Nature. 2011;473(7346):174–13. doi:10.1038/nature09944.
  • Faith JJ, Guruge JL, Charbonneau M, Subramanian S, Seedorf H, Goodman AL, Clemente JC, Knight R, Heath AC, Leibel RL, et al. The long-term stability of the human gut microbiota. Science. 2013;341(6141):1237439. doi:10.1126/science.1237439.
  • Almeida A, Mitchell AL, Boland M, Forster SC, Gloor GB, Tarkowska A, Lawley TD, Finn RD. A new genomic blueprint of the human gut microbiota. Nature. 2019;568(7753):499–504. Published online. doi:10.1038/s41586-019-0965-1.
  • Martinson JNV, Pinkham NV, Peters GW, Cho H, Heng J, Rauch M, Broadaway SC, Walk ST. Rethinking gut microbiome residency and the Enterobacteriaceae in healthy human adults. ISME J. 2019;13(9):2306–2318. doi:10.1038/s41396-019-0435-7.
  • Yang J, Pu J, Lu S, Bai X, Wu Y, Jin D, Cheng Y, Zhang G, Zhu W, Luo X, et al. Species-level analysis of human gut microbiota with metataxonomics. Front Microbiol. 2020;11:2029. doi:10.3389/fmicb.2020.02029.
  • Scepanovic P, Hodel F, Mondot S, Partula V, Byrd A, Hammer C, Alanio C, Bergstedt J, Patin E, Touvier M, et al. A comprehensive assessment of demographic, environmental, and host genetic associations with gut microbiome diversity in healthy individuals. Microbiome. 2019;7(1):130. doi:10.1186/s40168-019-0747-x.
  • Hansen MEB, Rubel MA, Bailey AG, Ranciaro A, Thompson SR, Campbell MC, Beggs W, Dave JR, Mokone GG, Mpoloka SW, et al. Population structure of human gut bacteria in a diverse cohort from rural Tanzania and Botswana. Genome Biol. 2019;20(1):16. doi:10.1186/s13059-018-1616-9.
  • Pires ES, Hardoim CCP, Miranda KR, Secco DA, Lobo LA, De Carvalho DP, Han J, Borchers CH, Ferreira RBR, Salles JF, et al. The gut microbiome and metabolome of two riparian communities in the Amazon. Front Microbiol. 2019;10:2003. doi:10.3389/fmicb.2019.02003.
  • Vangay P, Johnson AJ, Ward TL, Al-Ghalith GA, Shields-Cutler RR, Hillmann BM, Lucas SK, Beura LK, Thompson EA, Till LM, et al. US immigration westernizes the human gut microbiome. Cell. 2018;175(4):962–972.e10. doi:10.1016/j.cell.2018.10.029.
  • Schloig S, Arumugam M, Sunagawa S, Mitreva M, Tap J, Zhu A, Waller A, Mende DR, Kultima JR, Martin J, et al. Genomic variation landscape of the human gut microbiome. Nature. 2013;493(7430):45–50. doi:10.1038/nature11711.
  • Martin AM, Sun EW, Rogers GB, Keating DJ. The influence of the gut microbiome on host metabolism through the regulation of gut hormone release. Front Physiol. 2019;10(MAR). doi:10.3389/fphys.2019.00428.
  • Krishnan S, Alden N, Lee K. Pathways and functions of gut microbiota metabolism impacting host physiology. Curr Opin Biotechnol. 2015;36:137–145. doi:10.1016/j.copbio.2015.08.015.
  • Rüb AM, Tsakmaklis A, Gräfe SK, Simon MC, Vehreschild MJ, Wuethrich I. Biomarkers of human gut microbiota diversity and dysbiosis. Biomark Med. 2021;15(2):137–148. doi:10.2217/bmm-2020-0353.
  • Tamburini S, Shen N, Wu HC, Clemente JC. The microbiome in early life: implications for health outcomes. Nat Med. 2016;22(7):713–722. doi:10.1038/nm.4142.
  • Agustí A, García-Pardo MP, López-Almela I, Campillo I, Maes M, Romaní-Pérez M, Sanz Y. Interplay between the gut-brain axis, obesity and cognitive function. Front Neurosci. 2018;12(MAR):155. doi:10.3389/fnins.2018.00155.
  • Bibbò S, Ianiro G, Dore MP, Simonelli C, Newton EE, Cammarota G. Gut microbiota as a driver of inflammation in nonalcoholic fatty liver disease. Mediators Inflamm. 2018;2018:1–7. doi:10.1155/2018/9321643.
  • Tweedle JL, Deepe GS. Tumor necrosis factor alpha antagonism reveals a gut/lung axis that amplifies regulatory T cells in a pulmonary fungal infection. Infect Immun. 2018;86(6). doi:10.1128/IAI.00109-18.
  • Caballero S, Kim S, Carter RA, Leiner IM, Sušac B, Miller L, Kim GJ, Ling L, Pamer EG. Cooperating commensals restore colonization resistance to vancomycin-resistant Enterococcus faecium. Cell Host Microbe. 2017;21(5):592–602.e4. doi:10.1016/j.chom.2017.04.002.
  • Haak BW, Prescott HC, Wiersinga WJ. Therapeutic potential of the gut microbiota in the prevention and treatment of sepsis. Front Immunol. 2018;9(SEP):2042. doi:10.3389/fimmu.2018.02042.
  • Pickard JM, Zeng MY, Caruso R, Núñez G. Gut microbiota: role in pathogen colonization, immune responses, and inflammatory disease. Immunol Rev. 2017;279(1):70–89. doi:10.1111/imr.12567.
  • Kowalska-Duplaga K, Gosiewski T, Kapusta P, Sroka-Oleksiak A, Wędrychowicz A, Pieczarkowski S, Ludwig-Słomczyńska AH, Wołkow PP, Fyderek K. Differences in the intestinal microbiome of healthy children and patients with newly diagnosed Crohn’s disease. Sci Rep. 2019;9(1). doi:10.1038/s41598-019-55290-9.
  • Isaac S, Scher JU, Djukovic A, Jiménez N, Littman DR, Abramson SB, Pamer EG, Ubeda C. Short- and long-term effects of oral vancomycin on the human intestinal microbiota. J Antimicrob Chemother. 2017;72(1):128–136. doi:10.1093/jac/dkw383.
  • Sorbara MT, Pamer EG. Interbacterial mechanisms of colonization resistance and the strategies pathogens use to overcome them. Mucosal Immunol. 2019;12(1):1–9. doi:10.1038/s41385-018-0053-0.
  • Browne PD, Claassen E, Cabana MD. Microbiota in health and disease: from pregnancy to childhood. Benef Microbes. 2018;9(1):173–174. doi:10.3920/bm2017.x003.
  • Ciabattini A, Olivieri R, Lazzeri E, Medaglini D. Role of the microbiota in the modulation of vaccine immune responses. Front Microbiol. 2019;10:1305. doi:10.3389/FMICB.2019.01305.
  • Rakoff-Nahoum S, Paglino J, Eslami-Varzaneh F, Edberg S, Medzhitov R. Recognition of commensal microflora by toll-like receptors is required for intestinal homeostasis. Cell. 2004;118(2):229–241. doi:10.1016/j.cell.2004.07.002.
  • Kim S, Covington A, Pamer EG. The intestinal microbiota: antibiotics, colonization resistance, and enteric pathogens. Immunol Rev. 2017;279(1):90–105. doi:10.1111/imr.12563.
  • Takeda K, Kaisho T, Akira S. Toll-like receptors. Annu Rev Immunol. 2003;21(1):335–376. doi:10.1146/annurev.immunol.21.120601.141126.
  • Tartey S, Takeuchi O. Pathogen recognition and toll-like receptor targeted therapeutics in innate immune cells. Int Rev Immunol. 2017;36(2):57–73. doi:10.1080/08830185.2016.1261318.
  • Franchi L, Warner N, Viani K, Nuñez G. Function of nod-like receptors in microbial recognition and host defense. Immunol Rev. 2009;227(1):106–128. doi:10.1111/j.1600-065X.2008.00734.x.
  • Meunier E, Broz P. Evolutionary convergence and divergence in NLR function and structure. Trends Immunol. 2017;38(10):744–757. doi:10.1016/j.it.2017.04.005.
  • Yao X, Zhang C, Xing Y, Xue G, Zhang Q, Pan F, Wu G, Hu Y, Guo Q, Lu A, et al. Remodelling of the gut microbiota by hyperactive NLRP3 induces regulatory T cells to maintain homeostasis. Nat Commun. 2017;8(1):1896. doi:10.1038/s41467-017-01917-2.
  • Gewirtz AT, Navas TA, Lyons S, Godowski PJ, Madara JL. Cutting edge: bacterial flagellin activates basolaterally expressed TLR5 to induce epithelial proinflammatory gene expression. J Immunol. 2001;167(4):1882–1885. doi:10.4049/jimmunol.167.4.1882.
  • Sansonetti P. Host-pathogen interactions: the seduction of molecular cross talk. Gut. 2002;50(Supplement 3):iii2–iii8. doi:10.1136/gut.50.suppl_3.iii2.
  • Wang G, Huang S, Wang Y, Cai S, Yu H, Liu H, Zeng X, Zhang G, Qiao S. Bridging intestinal immunity and gut microbiota by metabolites. Cell Mol Life Sci. 2019;1:3. doi:10.1007/s00018-019-03190-6.
  • Levy M, Thaiss CA, Zeevi D, Dohnalová L, Zilberman-Schapira G, Mahdi JA, David E, Savidor A, Korem T, Herzig Y, et al. Microbiota-modulated metabolites shape the intestinal microenvironment by regulating NLRP6 inflammasome signaling. Cell. 2015;163(6):1428–1443. doi:10.1016/j.cell.2015.10.048.
  • Böcker U, Böcker U, Nebe T, Herweck F, Holt L, Panja A, Jobin C, Rossol S, Sartor RB, Singer MV. Immunomodulatory effects of butyrate on IEC butyrate modulates intestinal epithelial cell-mediated neutrophil migration. Clin Exp Immunol. 2003;131(1):53–60. doi:10.1046/j.1365-2249.2003.02056.x.
  • Sonnenberg GF, Artis D. Innate lymphoid cells in the initiation, regulation and resolution of inflammation. Nat Med. 2015;21(7):698–708. doi:10.1038/nm.3892.
  • Wilck N, Matus MG, Kearney SM, Olesen SW, Forslund K, Bartolomaeus H, Haase S, Mähler A, Balogh A, Markó L, et al. Salt-responsive gut commensal modulates TH17 axis and disease. Nature. 2017;551(7682):585–589. doi:10.1038/nature24628.
  • Sakaguchi S, Yamaguchi T, Nomura T, Ono M. Regulatory T cells and immune tolerance. Cell. Cell Press. 2008;133(5):775–787. doi:10.1016/j.cell.2008.05.009.
  • Takenaka MC, Quintana FJ. Tolerogenic dendritic cells. Semin Immunopathol. 2017;39(2):113–120. doi:10.1007/s00281-016-0587-8.
  • Tanoue T, Atarashi K, Honda K. Development and maintenance of intestinal regulatory T cells. Nat Rev Immunol. 2016;16(5):295–309. doi:10.1038/nri.2016.36.
  • Bene K, Varga Z, Petrov VO, Boyko N, Rajnavolgyi E. Gut microbiota species can provoke both inflammatory and tolerogenic immune responses in human dendritic cells mediated by retinoic acid receptor alpha ligation. Front Immunol. 2017;8(APR):427. doi:10.3389/fimmu.2017.00427.
  • Pascal M, Perez-Gordo M, Caballero T, Escribese MM, Lopez LMN, Luengo O, Manso L, Matheu V, Seoane E, Zamorano M, et al. Microbiome and allergic diseases. Front Immunol. 2018;9(JUL):1584. doi:10.3389/fimmu.2018.01584.
  • Muscogiuri G, Cantone E, Cassarano S, Tuccinardi D, Barrea L, Savastano S, Colao A. Gut microbiota: a new path to treat obesity. Int J Obes Suppl. 2019;9(1):10–19. doi:10.1038/s41367-019-0011-7.
  • Ahmadmehrabi S, Tang WHW. Gut microbiome and its role in cardiovascular diseases. Curr Opin Cardiol. 2017;32(6):761–766. doi:10.1097/HCO.0000000000000445.
  • Schwabe RF, Jobin C. The microbiome and cancer. Nat Rev Cancer. 2013;13(11):800–812. doi:10.1038/nrc3610.
  • Taur Y, Xavier JB, Lipuma L, Ubeda C, Goldberg J, Gobourne A, Lee YJ, Dubin KA, Socci ND, Viale A, et al. Intestinal domination and the risk of bacteremia in patients undergoing allogeneic hematopoietic stem cell transplantation. Clin Infect Dis. 2012;55(7):905–914. doi:10.1093/cid/cis580.
  • Shin JH, Warren CA. Prevention and treatment of recurrent Clostridioides difficile infection. Curr Opin Infect Dis. 2019;32(5):482–489. doi:10.1097/qco.0000000000000587.
  • Ianiro G, Maida M, Burisch J, Simonelli C, Hold G, Ventimiglia M, Gasbarrini A, Cammarota G. Efficacy of different faecal microbiota transplantation protocols for Clostridium difficile infection: a systematic review and meta-analysis. United Eur Gastroenterol J. 2018;6(8):1232–1244. doi:10.1177/2050640618780762.
  • Sorg JA, Sonenshein AL. Bile salts and glycine as cogerminants for Clostridium difficile spores. J Bacteriol. 2008;190(7):2505–2512. doi:10.1128/JB.01765-07.
  • Theriot CM, Koenigsknecht MJ, Carlson PE, Hatton GE, Nelson AM, Li B, Huffnagle GB, Li JZ, Young VB. Antibiotic-induced shifts in the mouse gut microbiome and metabolome increase susceptibility to Clostridium difficile infection. Nat Commun. 2014;5(1). doi:10.1038/ncomms4114.
  • Ng SC, Kamm MA, Yeoh YK, Chan PKS, Zuo T, Tang W, Sood A, Andoh A, Ohmiya N, Zhou Y, et al. Scientific frontiers in faecal microbiota transplantation: joint document of Asia-Pacific Association of Gastroenterology (APAGE) and Asia-Pacific Society for Digestive Endoscopy (APSDE). Gut. 2019. Published online. doi:10.1136/gutjnl-2019-319407.
  • Khoruts A. Is fecal microbiota transplantation a temporary patch for treatment of Clostridium difficile infection or a new frontier of therapeutics? Expert Rev Gastroenterol Hepatol. 2018;12(5):435–438. doi:10.1080/17474124.2018.1465818.
  • Cammarota G, Ianiro G, Tilg H, Rajilić-Stojanović M, Kump P, Satokari R, Sokol H, Arkkila P, Pintus C, Hart A, et al. European consensus conference on faecal microbiota transplantation in clinical practice. Gut. 2017;66(4):569–580. BMJ Publishing Group. doi:10.1136/gutjnl-2016-313017.
  • DeFilipp Z, Bloom PP, Torres Soto M, Mansour MK, Sater MRA, Huntley MH, Turbett S, Chung RT, Chen Y-B, Hohmann EL. Drug-resistant E. coli bacteremia transmitted by fecal microbiota transplant. N Engl J Med. 2019;381(21):2043–2050. doi:10.1056/NEJMoa1910437.
  • Jouhten H, Mattila E, Arkkila P, Satokari R. Reduction of antibiotic resistance genes in intestinal microbiota of patients with recurrent Clostridium difficile infection after fecal microbiota transplantation. Clin Infect Dis. 2016;63(5):710–711. doi:10.1093/cid/ciw390.
  • Leung V, Vincent C, Edens TJ, Miller M, Manges AR. Antimicrobial resistance gene acquisition and depletion following fecal microbiota transplantation for recurrent Clostridium difficile infection. Clin Infect Dis. 2018;66(3):456–457. doi:10.1093/cid/cix821.
  • Gopalsamy SN, Woodworth MH, Wang T, Carpentieri CT, Mehta N, Friedman-Moraco RJ, Mehta AK, Larsen CP, Kraft CS. The use of microbiome restoration therapeutics to eliminate intestinal colonization with multidrug-resistant organisms. Am J Med Sci. 2018;356(5):433–440. doi:10.1016/j.amjms.2018.08.015.
  • Dubberke ER, Mullane KM, Gerding DN, Lee CH, Louie TJ, Guthertz H, Jones C. Clearance of vancomycin-resistant Enterococcus concomitant with administration of a microbiota-based drug targeted at recurrent Clostridium difficile infection. Open Forum Infect Dis. 2016;3(3):ofw133. doi:10.1093/ofid/ofw133.
  • Davido B, Batista R, Fessi H, Salomon J, Dinh A. Impact of faecal microbiota transplantation to eradicate vancomycin-resistant enterococci (VRE) colonization in humans. J Infect. 2017;75(4):376–377. doi:10.1016/j.jinf.2017.06.001.
  • Kuijper EJ, Vendrik KEW, Vehreschild MJGT. Manipulation of the microbiota to eradicate multidrug-resistant Enterobacteriaceae from the human intestinal tract. Clin Microbiol Infect. 2019;25(7):786–789. doi:10.1016/j.cmi.2019.03.025.
  • Alagna L, Palomba E, Mangioni D, Bozzi G, Lombardi A, Ungaro R, Castelli V, Prati D, Vecchi M, Muscatello A, et al. Multidrug‐resistant gram‐negative bacteria decolonization in immunocompromised patients: a focus on fecal microbiota transplantation. Int J Mol Sci. 2020;21(16):1–22. doi:10.3390/ijms21165619.
  • Ueckermann V, Hoosien E, De Villiers N, Geldenhuys J. Fecal microbial transplantation for the treatment of persistent multidrug-resistant Klebsiella pneumoniae infection in a critically ill patient. Case Rep Infect Dis. 2020;2020:1–5. doi:10.1155/2020/8462659.
  • Gouveia C, Palos C, Pereira P, Roque Ramos L, Cravo M. Fecal microbiota transplant in a patient infected with multidrug-resistant bacteria: a case report. GE Port J Gastroenterol. 2021;28(1):56–61. doi:10.1159/000507263.
  • Allegretti JR, Mullish BH, Kelly C, Fischer M. The evolution of the use of faecal microbiota transplantation and emerging therapeutic indications. Lancet. 2019;394(10196):420–431. doi:10.1016/S0140-6736(19)31266-8.
  • Balakrishnan B, Taneja V. Microbial modulation of the gut microbiome for treating autoimmune diseases. Expert Rev Gastroenterol Hepatol. 2018;12(10):985–996. doi:10.1080/17474124.2018.1517044.
  • Markowiak P, Ślizewska K. Effects of probiotics, prebiotics, and synbiotics on human health. Nutrients. 2017;9(9). doi:10.3390/nu9091021.
  • O’Toole PW, Marchesi JR, Hill C. Next-generation probiotics: the spectrum from probiotics to live biotherapeutics. Nat Microbiol. 2017;2(5). doi:10.1038/nmicrobiol.2017.57.
  • Zipperer A, Konnerth MC, Laux C, Berscheid A, Janek D, Weidenmaier C, Burian M, Schilling NA, Slavetinsky C, Marschal M, et al. Human commensals producing a novel antibiotic impair pathogen colonization. Nature. 2016;535(7613):511–516. doi:10.1038/nature18634.
  • Bitschar K, Sauer B, Focken J, Dehmer H, Moos S, Konnerth M, Schilling NA, Grond S, Kalbacher H, Kurschus FC, et al. Lugdunin amplifies innate immune responses in the skin in synergy with host- and microbiota-derived factors. Nat Commun. 2019;10(1). doi:10.1038/s41467-019-10646-7.
  • Sorbara MT, Dubin K, Littmann ER, Moody TU, Fontana E, Seok R, Leiner IM, Taur Y, Peled JU, Van Den Brink MRM, et al. Inhibiting antibiotic-resistant Enterobacteriaceae by microbiota-mediated intracellular acidification. J Exp Med. 2019;216(1):84–98. doi:10.1084/JEM.20181639.
  • Imperial ICVJ, Ibana JA. Addressing the antibiotic resistance problem with probiotics: reducing the risk of its double-edged sword effect. Front Microbiol. 2016;7(DEC):1983. doi:10.3389/fmicb.2016.01983.
  • Manley KJ, Fraenkel MB, Mayall BC, Power DA. Probiotic treatment of vancomycin-resistant enterococci: a randomised controlled trial. Med J Aust. 2007;186(9):454–457. doi:10.5694/j.1326-5377.2007.tb00995.x.
  • Szachta P, Ignyś I, Cichy W. An evaluation of the ability of the probiotic strain Lactobacillus rhamnosus GG to eliminate the gastrointestinal carrier state of vancomycin-resistant Enterococci in colonized children. J Clin Gastroenterol. 2011;45(10):872–877. doi:10.1097/MCG.0b013e318227439f.
  • Salomão MCC, Heluany-Filho MA, Menegueti MG, De Kraker MEA, Martinez R, Bellissimo-Rodrigues F. A randomized clinical trial on the effectiveness of a symbiotic product to decolonize patients harboring multidrug-resistant gram-negative bacilli. Rev Soc Bras Med Trop. 2016;49(5):559–566. doi:10.1590/0037-8682-0233-2016.
  • Aguilar-Toalá JE, Garcia-Varela R, Garcia HS, Mata-Haro V, González-Córdova AF, Vallejo-Cordoba B, Hernández-Mendoza A. Postbiotics: an evolving term within the functional foods field. Trends Food Sci Technol. 2018;75:105–114. doi:10.1016/j.tifs.2018.03.009.
  • Clifford MN. Diet-derived phenols in plasma and tissues and their implications for health. Planta Med. 2004;70(12):1103–1114. doi:10.1055/s-2004-835835.
  • Gibson GR, Hutkins R, Sanders ME, Prescott SL, Reimer RA, Salminen SJ, Scott K, Stanton C, Swanson KS, Cani PD, et al. Expert consensus document: the International Scientific Association for Probiotics and Prebiotics (ISAPP) consensus statement on the definition and scope of prebiotics. Nat Rev Gastroenterol Hepatol. 2017;14(8):491–502. doi:10.1038/nrgastro.2017.75.
  • Dueñas M, Muñoz-González I, Cueva C, Jiménez-Girón A, Sánchez-Patán F, Santos-Buelga C, Moreno-Arribas MV, Bartolomé B. A survey of modulation of gut microbiota by dietary polyphenols. Biomed Res Int. 2015:2015. doi:10.1155/2015/850902.
  • Elison E, Vigsnaes LK, Rindom Krogsgaard L, Rasmussen J, Sorensen N, McConnell B, Hennet T, Sommer MOA, Bytzer P. Oral supplementation of healthy adults with 2′-O-fucosyllactose and lacto-N-neotetraose is well tolerated and shifts the intestinal microbiota. Br J Nutr. 2016;116(8):1356–1368. doi:10.1017/S0007114516003354.
  • Kulinich A, Liu L. Human milk oligosaccharides: the role in the fine-tuning of innate immune responses. Carbohydr Res. 2016;432:62–70. doi:10.1016/j.carres.2016.07.009.
  • Morrow AL, Ruiz-Palacios GM, Jiang X, Newburg DS. Human-milk glycans that inhibit pathogen binding protect breast-feeding infants against infectious diarrhea. J Nutr. 2005;135(5):1304–1307. doi:10.1093/jn/135.5.1304.
  • Fábrega MJ, Aguilera L, Giménez R, Varela E, Cañas MA, Antolín M, Badía J, Baldomà L. Activation of immune and defense responses in the intestinal mucosa by outer membrane vesicles of commensal and probiotic Escherichia coli strains. Front Microbiol. 2016;7(MAY). doi:10.3389/fmicb.2016.00705.
  • Chelakkot C, Choi Y, Kim DK, Park HT, Ghim J, Kwon Y, Jeon J, Kim MS, Jee YK, Gho YS, et al. Akkermansia muciniphila-derived extracellular vesicles influence gut permeability through the regulation of tight junctions. Exp Mol Med. 2018;50(2). doi:10.1038/emm.2017.282.
  • Lawrence, Baldridge, Handley. Phages and human health: more than idle hitchhikers. Viruses. 2019;11(7):587. doi:10.3390/v11070587.
  • Pal C, Maciá MD, Oliver A, Schachar I, Buckling A. Coevolution with viruses drives the evolution of bacterial mutation rates. Nature. 2007;450(7172):1079–1081. doi:10.1038/nature06350.
  • Enault F, Briet A, Bouteille L, Roux S, Sullivan MB, Petit MA. Phages rarely encode antibiotic resistance genes: a cautionary tale for virome analyses. Isme J. 2017;11(1):237–247. doi:10.1038/ismej.2016.90.
  • Modi SR, Lee HH, Spina CS, Collins JJ. Antibiotic treatment expands the resistance reservoir and ecological network of the phage metagenome. Nature. 2013;499(7457):219–222. doi:10.1038/nature12212.
  • Torres-Barceló C. The disparate effects of bacteriophages on antibiotic-resistant bacteria. Emerg Microbes Infect. 2018;7(1):1–12. doi:10.1038/s41426-018-0169-z.
  • D’Herelle F. On an invisible microbe antagonistic to dysentery bacilli. Comptes Rendus Acad des Sci. 1917;165:373–375. doi:10.4161/bact.1.1.14941.
  • D’Herelle F, Malone RH. A preliminary report of work carried out by the cholera bacteriophage enquiry. Ind Med Gaz. 1927;62:614–616.
  • Kakasis A, Panitsa G. Bacteriophage therapy as an alternative treatment for human infections. A comprehensive review. Int J Antimicrob Agents. 2019;53(1):16–21. doi:10.1016/j.ijantimicag.2018.09.004.
  • El Haddad L, Harb CP, Gebara MA, Stibich MA, Chemaly RF. A systematic and critical review of bacteriophage therapy against multidrug-resistant ESKAPE organisms in humans. Clin Infect Dis. 2019;69(1):167–178. doi:10.1093/cid/ciy947.
  • Gordillo Altamirano FL, Barr JJ. Phage therapy in the postantibiotic era. Clin Microbiol Rev. 2019;32(2). doi:10.1128/CMR.00066-18.
  • Ross A, Ward S, Hyman P. More is better: selecting for broad host range bacteriophages. Front Microbiol. 2016;7(SEP):1352. doi:10.3389/fmicb.2016.01352.
  • Biswas B, Adhya S, Washart P, Paul B, Trostel AN, Powell B, Carlton R, Merril CR. Bacteriophage therapy rescues mice bacteremic from a clinical isolate of vancomycin-resistant Enterococcus faecium. Infect Immun. 2002;70(1):204–210. doi:10.1128/IAI.70.1.204-210.2002.
  • Abd El-Aziz AM, Elgaml A, Ali YM. Bacteriophage therapy increases complement-mediated lysis of bacteria and enhances bacterial clearance after acute lung infection with multidrug-resistant Pseudomonas aeruginosa. J Infect Dis. 2019;219(9):1439–1447. doi:10.1093/infdis/jiy678.
  • Hua Y, Luo T, Yang Y, Dong D, Wang R, Wang Y, Xu M, Guo X, Hu F, He P. Phage therapy as a promising new treatment for lung infection caused by carbapenem-resistant Acinetobacter baumannii in mice. Front Microbiol. 2018;8:2659. doi:10.3389/fmicb.2017.02659.
  • Wright A, Hawkins CH, Änggård EE, Harper DR. A controlled clinical trial of a therapeutic bacteriophage preparation in chronic otitis due to antibiotic-resistant Pseudomonas aeruginosa; a preliminary report of efficacy. Clin Otolaryngol. 2009;34(4):349–357. doi:10.1111/j.1749-4486.2009.01973.x.
  • Rhoads DD, Wolcott RD, Kuskowski MA, Wolcott BM, Ward LS, Sulakvelidze A. Bacteriophage therapy of venous leg ulcers in humans: results of a phase I safety trial. J Wound Care. 2009;18(6):237–243. doi:10.12968/jowc.2009.18.6.42801.
  • Rose T, Verbeken G, De VD, Merabishvili M, Vaneechoutte M, Lavigne R, Jennes S, Zizi M, Pirnay J-P. Experimental phage therapy of burn wound infection: difficult first steps. Int J Burns Trauma. 2014;4(2):66–73. PMID: 25356373.
  • Sarker SA, Sultana S, Reuteler G, Moine D, Descombes P, Charton F, Bourdin G, McCallin S, Ngom-Bru C, Neville T, et al. Oral phage therapy of acute bacterial diarrhea with two coliphage preparations: a randomized trial in children from Bangladesh. EBioMedicine. 2016;4:124–137. doi:10.1016/j.ebiom.2015.12.023.
  • Ooi ML, Drilling AJ, Morales S, Fong S, Moraitis S, MacIas-Valle L, Vreugde S, Psaltis AJ, Wormald PJ. Safety and tolerability of bacteriophage therapy for chronic rhinosinusitis due to Staphylococcus aureus. JAMA Otolaryngol - Head Neck Surg. 2019;145(8):723. doi:10.1001/jamaoto.2019.1191.
  • Duplessis C, Biswas B, Hanisch B, Perkins M, Henry M, Quinones J, Wolfe D, Estrella L, Hamilton T. Refractory Pseudomonas bacteremia in a 2-year-old sterilized by bacteriophage therapy. J Pediatric Infect Dis Soc. 2018;7(3):253–256. doi:10.1093/jpids/pix056.
  • Karch CP, Burkhard P. Vaccine technologies: from whole organisms to rationally designed protein assemblies. Biochem Pharmacol. 2016;120:1–14. doi:10.1016/j.bcp.2016.05.001.
  • Moyer TJ, Zmolek AC, Irvine DJ. Beyond antigens and adjuvants: formulating future vaccines. J Clin Invest. 2016;126(3):799–808. doi:10.1172/JCI81083.
  • Tiller KE, Tessier PM. Advances in antibody design. Annu Rev Biomed Eng. 2015;17(1):191–216. doi:10.1146/annurev-bioeng-071114-040733.
  • Jansen KU, Knirsch C, Anderson AS. The role of vaccines in preventing bacterial antimicrobial resistance. Nat Med. 2018;24(1):10–20. doi:10.1038/nm.4465.
  • Lipsitch M, Siber GR. How can vaccines contribute to solving the antimicrobial resistance problem? MBio. 2016;7(3). doi:10.1128/mBio.00428-16.
  • Chang -H-H, Cohen T, Grad YH, Hanage WP, O’Brien TF, Lipsitch M. Origin and proliferation of multiple-drug resistance in bacterial pathogens. Microbiol Mol Biol Rev. 2015;79(1):101–116. doi:10.1128/mmbr.00039-14.
  • Kennedy DA, Read AF. Why the evolution of vaccine resistance is less of a concern than the evolution of drug resistance. Proc Natl Acad Sci USA. 2018;115(51):12878–12886. doi:10.1073/pnas.1717159115.
  • Kawaguchiya M, Urushibara N, Aung MS, Shinagawa M, Takahashi S, Kobayashi N. Serotype distribution, antimicrobial resistance and prevalence of pilus islets in pneumococci following the use of conjugate vaccines. J Med Microbiol. 2017;66(5):643–650. doi:10.1099/jmm.0.000479.
  • Cerquetti M, Giufrè M. Why we need a vaccine for non-typeable Haemophilus influenzae. Hum Vaccines Immunother. 2016;12(9):2357–2361. doi:10.1080/21645515.2016.1174354.
  • Langereis JD, De Jonge MI. Invasive disease caused by nontypeable Haemophilus influenzae. Emerg Infect Dis. 2015;21(10). doi:10.3201/eid2110.150004.
  • Van Eldere J, Slack MPE, Ladhani S, Cripps AW. Non-typeable Haemophilus influenzae, an under-recognised pathogen. Lancet Infect Dis. 2014;14(12):1281–1292. doi:10.1016/S1473-3099(14)70734-0.
  • Kapatai G, Sheppard CL, Al-Shahib A, Litt DJ, Underwood AP, Harrison TG, Fry NK. Whole genome sequencing of Streptococcus pneumoniae: development, evaluation and verification of targets for serogroup and serotype prediction using an automated pipeline. PeerJ. 2016;4:e2477. doi:10.7717/peerj.2477.
  • Juhas M. Horizontal gene transfer in human pathogens. Crit Rev Microbiol. 2015;41(1):101–108. doi:10.3109/1040841X.2013.804031.
  • Cooper RM, Tsimring L, Hasty J. Inter-species population dynamics enhance microbial horizontal gene transfer and spread of antibiotic resistance. Elife. 2017:6. doi:10.7554/eLife.25950.
  • Senna JP, Roth DM, Oliveira JS, Machado DC, Santos DS. Protective immune response against methicillin resistant Staphylococcus aureus in a murine model using a DNA vaccine approach. Vaccine. 2003;21(19–20):2661–2666. doi:10.1016/S0264-410X(02)00738-7.
  • Ciofu O, Bagge N, Hoiby N. Antibodies against beta-lactamase can improve ceftazidime treatment of lung infection with beta-lactam-resistant Pseudomonas aeruginosa in a rat model of chronic lung infection. APMIS. 2002;110(12):881–891. doi:10.1034/j.1600-0463.2002.1101207.x.
  • Oh JZ, Ravindran R, Chassaing B, Carvalho FA, Maddur MS, Bower M, Hakimpour P, Gill KP, Nakaya HI, Yarovinsky F, et al. TLR5-mediated sensing of gut microbiota is necessary for antibody responses to seasonal influenza vaccination. Immunity. 2014;41(3):478–492. doi:10.1016/j.immuni.2014.08.009.
  • Kinnebrew MA, Buffie CG, Diehl GE, Zenewicz LA, Leiner I, Hohl TM, Flavell RA, Littman DR, Pamer EG. Interleukin 23 production by intestinal CD103+ CD11b+ dendritic cells in response to bacterial flagellin enhances mucosal innate immune defense. Immunity. 2012;36(2):276–287. doi:10.1016/j.immuni.2011.12.011.
  • Kim D, Kim YG, Seo SU, Kim DJ, Kamada N, Prescott D, Philpott DJ, Rosenstiel P, Inohara N, Núñez G. Nod2-mediated recognition of the microbiota is critical for mucosal adjuvant activity of cholera toxin. Nat Med. 2016;22(5):524–530. doi:10.1038/nm.4075.
  • Duthie MS, Windish HP, Fox CB, Reed SG. Use of defined TLR ligands as adjuvants within human vaccines. Immunol Rev. 2011;239(1):178–196. doi:10.1111/j.1600-065X.2010.00978.x.
  • Willemsen I, Groenhuijzen A, Bogaers D, Stuurman A, Van Keulen P, Kluytmans J. Appropriateness of antimicrobial therapy measured by repeated prevalence surveys. Antimicrob Agents Chemother. 2007;51(3):864–867. doi:10.1128/AAC.00994-06.
  • Davey P, Marwick CA, Scott CL, Charani E, Mcneil K, Brown E, Gould IM, Ramsay CR, Michie S. Interventions to improve antibiotic prescribing practices for hospital inpatients. Cochrane Database Syst Rev. 2017;2017(2):CD003543. doi:10.1002/14651858.CD003543.pub4.
  • Bhalodi AA, Van Engelen TSR, Virk HS, Wiersinga WJ. Impact of antimicrobial therapy on the gut microbiome. J Antimicrob Chemother. 2019;74(Suppl 1):I6–I15. doi:10.1093/jac/dky530.
  • Wipperman MF, Fitzgerald DW, Juste MAJ, Taur Y, Namasivayam S, Sher A, Bean JM, Bucci V, Glickman MS. Antibiotic treatment for tuberculosis induces a profound dysbiosis of the microbiome that persists long after therapy is completed. Sci Rep. 2017;7(1):1–11. doi:10.1038/s41598-017-10346-6.
  • Neuman H, Forsythe P, Uzan A, Avni O, Koren O. Antibiotics in early life: dysbiosis and the damage done. FEMS Microbiol Rev. 2018;42(4):489–499. doi:10.1093/femsre/fuy018.
  • Qian X, Yanagi K, Kane AV, Alden N, Lei M, Snydman DR, Vickers RJ, Lee K, Thorpe CM. Ridinilazole, a narrow spectrum antibiotic for treatment of Clostridioides difficile infection, enhances preservation of microbiota-dependent bile acids. Am J Physiol - Gastrointest Liver Physiol. 2020;319(2):G227–G237. doi:10.1152/ajpgi.00046.2020.
  • Mariathasan S, Tan MW. Antibody–antibiotic conjugates: a novel therapeutic platform against bacterial infections. Trends Mol Med. 2017;23(2):135–149. doi:10.1016/j.molmed.2016.12.008.
  • Gao W, Chen Y, Zhang Y, Zhang Q, Zhang L. Nanoparticle-based local antimicrobial drug delivery. Adv Drug Deliv Rev. 2018;127:46–57. doi:10.1016/j.addr.2017.09.015.
  • Connelly S, Fanelli B, Hasan NA, Colwell RR, Kaleko M. Oral metallo-beta-lactamase protects the gut microbiome from carbapenem-mediated damage and reduces propagation of antibiotic resistance in pigs. Front Microbiol. 2019;10(FEB):101. doi:10.3389/fmicb.2019.00101.
  • Septimus EJ. Antimicrobial resistance: an antimicrobial/diagnostic stewardship and infection prevention approach. Med Clin North Am. 2018;102(5):819–829. doi:10.1016/j.mcna.2018.04.005.
  • De Jong E, Van Oers JA, Beishuizen A, Vos P, Vermeijden WJ, Haas LE, Loef BG, Dormans T, Van Melsen GC, Kluiters YC, et al. Efficacy and safety of procalcitonin guidance in reducing the duration of antibiotic treatment in critically ill patients: a randomised, controlled, open-label trial. Lancet Infect Dis. 2016;16(7):819–827. doi:10.1016/S1473-3099(16)00053-0.
  • Baggs J, Jernigan JA, Halpin AL, Epstein L, Hatfield KM, McDonald LC. Risk of subsequent sepsis within 90 days after a hospital stay by type of antibiotic exposure. Clin Infect Dis. 2018;66(7):1004–1012. doi:10.1093/cid/cix947.
  • Baur D, Gladstone BP, Burkert F, Carrara E, Foschi F, Döbele S, Tacconelli E. Effect of antibiotic stewardship on the incidence of infection and colonisation with antibiotic-resistant bacteria and Clostridium difficile infection: a systematic review and meta-analysis. Lancet Infect Dis. 2017;17(9):990–1001. doi:10.1016/S1473-3099(17)30325-0.