Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2017 Nov 17.
Published in final edited form as: Microbiol Spectr. 2014 Nov;5(6):10.1128/microbiolspec.BAD-0014-2016. doi: 10.1128/microbiolspec.BAD-0014-2016

Enterococci and their interactions with the intestinal microbiome

Krista Dubin 1,2,*, Eric G Pamer 1,2,3
PMCID: PMC5691600  NIHMSID: NIHMS914371  PMID: 29125098

BACKGROUND

The Enterococcus genus is comprised of over 50 species that can be found in diverse environments, from the soil to the gastrointestinal (GI) tract of animals and humans to the hospital environment (1, 2). The first member of this gram-positive genus was isolated in 1899 from a lethal case of endocarditis (3, 4). It was not until 1984 that enterococcal species were seen as genetically distinct from Streptococcus and assigned their own genus (3, 4, 5). Enterococci are gram-positive facultative anaerobes that exist in chains or pairs and do not form spores. They grow optimally at 35 °C, hydrolyze esculin in the presence of 40% bile salts, and are catalase negative (6, 7). Enterococcal species can be distinguished by phenotypic tests that rely on strains’ ability to form acid in mannitol and sorbose broth, and to hydrolyze arginine (8, 9).

Enterococci are found within the fecal content of insects, birds, reptiles, and mammals (2, 10). Named ‘Entero’ to denote their intestinal residence, Enterococcus faecalis and faecium were first isolated in the early 1900s (11, 12, 13). Based on SNPs within 16S ribosomal RNA (rRNA), Enterococci are divided into seven evolutionarily distinct groups (14). E. faecalis is found in a host of different animals, suggesting that it was in evolutionarily terms an early gut colonizer (14). In humans, E. faecalis and E. faecium are the most abundant species of this genus found in fecal content, comprising up to 1% of the adult intestinal microbiota (15, 16, 17, 18, 19).

Enterococci have recently emerged as a prevalent multidrug-resistant nosocomical pathogen. Since the late 1970s and 1980s, enterococcal species have developed increased resistance to several classes of antibiotics (20, 21, 14). Resistant Enterococci densely colonize the gut following antibiotic treatment, which can deplete the GI tract of large swaths of protective commensals (22, 23, 24). Antibiotic use has increased the spread of drug-resistant Enterococci within the hospital setting, leading to Enterococci becoming one of the most common causes of hospital-associated infections (25).

Restoration of the intestinal microbiota to a healthy state is a new and developing approach to counter the continuing emergence of antibiotic-resistant microorganisms. However, manipulating the intestinal microbiome to prevent the spread of antibiotic-resistant bacterial strains, while also supporting the sensitive ecosystem of which Enterococci are constituents, is a delicate task. It requires that we understand the relationship of Enterococci to their natural intestinal habitat in the context of Enterococci’s dual life as commensals and nosocomial pathogens. To do so, we discuss the road Enterococci have traveled to become multi-drug-resistant hospital-associated infectious agents that possess diversified genomes that allows them to survive in the post-antibiotic intestinal niche. With that in mind, we can consider how best to manipulate or restore the enteric microbiota to benefit human health. In this chapter, we discuss the Enterococci’s 1) clinical importance, 2) development of antibiotic resistance, 3) diversity in genomic composition and habitats, and 4) interaction with the intestinal microbiome that may help limit its infectious spread.

CLINICAL IMPORTANCE

Infections

Enterococci emerged as a leading hospital-associated pathogen in the late 1970s and 1980s (26). In the US, Enterococci cause roughly 66,000 infections each year (27). Enterococci are often cultured from mixed species infections of the pelvis, abdomen and other soft tissues (28). Although the role that Enterococci play in these infections is not often clear, they are frequently treated with antibiotics. Less commonly, Enterococci can cause meningitis and septic arthritis in patients with comorbidities or who are immunocompromised (28).

Even more clinically important, Enterococci are leading causes of hospital-associated bacteremia, endocarditis and urinary tract infections (UTIs) (20, 26, 29). Enterococci are the second most common cause of nosocomial bacteremia and are associated with an overall mortality of roughly 33% (25, 30). Enterococcal bacteremia is often preceded by dense colonization of the GI tract, from which Enterococci can translocate into the bloodstream (31, 23). In addition, the loss of mucosal immunity and disruption of the GI barrier have been associated with enterococcal bacteremia; risk factors include mucositis, Clostridium difficile infection, and neutropenia (32, 33, 34).

Over 10% of infective endocarditis cases seen in North America are caused by Enterococci, making it the second leading cause (35). Of the total cases of enterococcal endocarditis, more than 35% of infections are acquired in the hospital (36). Enterococci form biofilms on damaged heart valves, which grow into structures called vegetations. Prosthetic values can also serve as a platform for enterococcal growth (36). As with bacteremia, Enterococci that cause endocarditis are often former inhabitants of the GI or genitourinary (GU) tract that gained access the bloodstream (37, 38). Over 10% of catheter-associated UTIs are of enterococcal origin (29).

Transmission and sources of infectious Enterococci

Hospital-acquired enterococcal infections are of particular concern due to both their increasing prevalence and growing resistance to antibiotics. Enterococci can readily spread within hospital units (39, 40, 41, 42, 43, 44). Transmission of Enterococci in the clinical environment is aided by two key factors: the ability of Enterococci to survive outside the GI tract, and the potential for healthcare workers to inadvertently transfer bacteria to adjacent patients. Enterococcal species can survive for prolonged periods on hospital surfaces, such as medical devices and bed rails, creating fomites that are a major risk factor for further spread (45, 46). Enterococci are transferred from patient to patient via healthcare workers’ hands (47, 48). Contaminated hands of medical staff can transfer vancomycin-resistant Enterococci (VRE) to roughly 1 out of every 10 clean surfaces that the healthcare workers touch (49).

The GI tract represents the major site colonized by antibiotic-resistant Enterococci and thus constitutes an important source of hospital-associated infections. Hospital contamination is increased when colonized patients become incontinent (50). The density of VRE in patients’ fecal content is correlated with the number of VRE transmission events (47). For roughly every 10% increase in patients colonized with VRE, the risk of additional hospitalized individuals acquiring VRE rises by 40% (46). A critical mechanism by which hospitalized patients become densely colonized with VRE is antibiotic treatment; how antibiotics allow for VRE expansion is detailed in the last sections of this chapter. The majority of antibiotic regimens with anti-anaerobic activity result in high-burden intestinal VRE density (22). Metronidazole increases the risk for high-density VRE colonization by 3-fold in allogeneic hematopoietic stem cell transplant (allo-HSCT) patient cohorts (24). Other risk factors for colonization include use of catheters in the bloodstream or urinary tract, prior surgery, length of hospital stay, and exposure to VRE-colonized patients (51, 52, 40, 53, 47, 54).

Treatment

Severe cases of enterococcal infections, such as infections of heart valves, has relied on combination drug therapy (55, 56). Combined administration of penicillin and streptomycin (a beta-lactam and aminoglycoside, respectively) successfully cured 80% of enterococcal infective endocarditis (IE) cases, which previously had a mortality rate of between 20% to 50%, and became standard therapy by the 1950s (38, 57). Today, for IE caused by ampicillin- and vancomycin-sensitive E. faecalis lacking high-level resistance to aminoglycosides, gentamicin is the preferred aminoglycoside used in combination with ampicillin. Ampicillin plus ceftriaxone is an alternative therapy for ampicillin-susceptible E. faecalis. This regime has also been used to treat aminoglycoside-sensitive E. faecalis isolates, as it is associated with similar cure rates and less nephrotoxicity compared to ampicillin-gentamicin therapy (58, 59). Although E. faecalis isolates are intrinsically resistant to cephalosporins, the two beta-lactam antibiotics work synergistically by binding different pencillin-binding proteins (PBPs), the enzymes involved in bacterial cell wall synthesis (60). For ampicillin- and vancomycin-resistant isolates causing IE, the majority of which are E. faecium, daptomycin or linezolid can be used, although the clinical data as to their efficacy is limited (61, 62).

There are few other examples of bactericidal synergy against Enterococci, and novel antibiotic therapies are urgently needed for multi-drug resistant species (37). This clinical picture begs the question: how did commensal Enterococci become such a challenging pathogen? The plasticity of the enterococcal genome is a key factor that has allowed the bacteria to 1) acquire traits that confer antibiotic resistance through mobile genetic elements, 2) diversify over time into lineages specifically adapted to the hospital environment, and 3) colonize the GI tract at greater densities following antibiotic exposure (37). We discuss each point in the following three sections.

DEVELOPMENT OF ANTIBIOTIC RESISTANCE

Development of antibiotic resistance in E. faecium and E. faecalis

Roughly one-third of enterococcal infections in the US are drug-resistant, totaling 20,000 antibiotic-resistant cases per year, from which an estimated 1,300 patients succumb yearly (27). E. faecalis caused over 90% of clinical infections until the mid-1990s, at which point E. faecium became more clinically prevalent (63, 64). The rise of nosocomial E. faecium strains has been attributed to the increased use of vancomycin and broad-spectrum antibiotics (20, 25, 37, 65). To date in the US, E. faecium causes nearly a third of all enterococcal nosocomial infections and constitutes over 75% of all healthcare-associated VRE strains (29, 27). The majority of E. faecium infections associated with medical equipment are vancomycin-resistant and ampicillin-resistant (80–87% and 90%, respectively) (25, 66).

VRE emerged in the mid-1980s, first in Europe among livestock and then in the US within hospitals (67, 68). In the US, glycopeptide resistance developed among hospital-adapted ampicillin-resistant isolates that were the predominant Enterococci within hospital intestinal microbiota (21, 65). Vancomycin-resistant isolates have been associated with oral vancomycin used to treat antibiotic-associated diarrhea due to C. difficile in hospitalized patients. Of note, administration of vancomycin intravenously (IV) is not correlated with the development of VRE infection (69, 22, 24). Vancomycin by this route results in low intestinal concentrations (70). In Europe, the issue of VRE was initially confined to animal husbandry. VRE was seen in livestock regularly exposed to antibiotics. Avoparcin, a growth-promoting antibiotic that also provides cross-resistance to vancomycin, is thought to have contributed to the rise of VRE (71, 72, 73). Avoparcin was subsequently banned from use in 1996, and the prevalence of VRE in animals decreased (74, 75, 76). However, VRE has made a recent appearance in European hospitals with isolates closely related to healthcare-associated strains found in the US (77).

Enterococci harbor resistance through two means: 1) resistance that is encoded in the core genome of all enterococcal strains (intrinsic), and 2) resistance that is passed among isolates on mobile genetic elements by horizontal transfer (acquired). An overview of some of the mechanisms by which Enterococci developed resistance to ampicillin, vancomycin and daptomycin are briefly outlined.

Antibiotic resistance: Ampicillin

Beta-lactams, such as ampicillin, inhibit bacterial growth by modifying and thereby inactivating a group of enzymes called penicillin-binding proteins (PBPs). PBPs cross-link side chains of peptidoglycan peptides during cell wall synthesis. Enterococcal strains harbor some intrinsic resistance to beta-lactams by producing penicillin-binding protein 5 (PBP5), which is chromosomally encoded (78, 79). Given their low affinity to beta-lactam drugs, PBP5s can continue peptidoglycan synthesis as other PBPs become modified (80). Increased resistance to ampicillin is associated with mutations to the PBP5-encoding gene that further reduce the protein’s affinity for beta-lactam antibiotics, such as mutations that result in amino acid substitutions near the active site (81, 82, 83). Resistance is further amplified when multiple mutations are present in the pbp5 gene (83). Mutated alleles can be horizontally transferred to beta-lactam susceptible strains in vitro (84). Altogether, the pbp5 gene differs in nucleotide sequence by about 5% between sensitive and resistant strains (85). The acquisition of specific pbp5 gene mutations contributed to the high-level ampicillin resistance that nosocomial E. faecium isolates developed in the late 1970s and 1980s (85, 21, 86).

Antibiotic resistance: Vancomycin

Glycopeptide antibiotics, such as vancomycin, prevent peptidoglycan cell wall synthesis by forming complexes with the D-Ala-D-Ala peptide terminus of peptidoglycan precursors, blocking enzymatic binding sites. Resistant isolates alter peptidoglycan precursors to form D-Ala-D-Lactate or D-Ala-D-Serine, with 1000-fold to 7-fold lower drug binding affinity respectively (87, 88, 65). These modifications inhibit antibiotic binding while still allowing PBP enzymes to use these substrates to build a functional cell wall. In Enterococci, 9 genes clusters associated with resistance have been identified, with most being encoded on mobile elements (65). In response to glycopeptides, these resistance operons regulate the expression of a suite of enzymes that together create modified peptidoglycan precursors and remove those that are unaltered. The two major resistance operons are VanA and VanB (88). VanA gene loci are encoded on Tn1546 or related transposons, conferring high-level resistance to vancomycin and teicoplanin. VanB gene clusters are found on Tn5382/Tn1549-type transposons either on plasmids on in the chromosome, providing moderate resistance to vancomycin only. Variants of these vancomycin resistance gene loci are found worldwide (89).

Antibiotic resistance: Daptomycin

Daptomycin is a recently introduced antibiotic for the treatment of multi-drug resistant Enterococci; however its bactericidal mechanism of action is not fully understood. It is thought to alter the cytoplasmic membrane and cause depolarization in a calcium-dependent manner, leading to a release of potassium ions from the cell and subsequent cell death (90, 91). For Enterococci, the ability to resist Daptomycin in part results from alterations in the composition of its cell membrane and envelope. Whole-genome sequencing of a pair of sequentially isolated vancomycin-resistant E. faecalis clones, the first daptomycin sensitive and the second resistant, from a single patient’s bloodstream identified in-frame deletions in three genes: cls, gdpD, and liaf (92). cls and gdpD encode proteins thought to play a role in phospholipid metabolism, and liaF is part of a regulatory system that coordinates the cell-envelope response to antibiotics. Resequencing experiments found resistance-associated mutations that became fixed after only two weeks of in vitro serial passage with increasing concentrations of daptomycin (93). The transfer of cls mutation to susceptible E. faecalis strains confers resistance to daptomycin (93). Comparative sequencing analyses were performed on 5 vancomycin-resistant E. faecium strain pairs, all initially susceptible and then later resistant to daptomycin, that colonized HSCT patients’ GI tracts (94). These intestinal VRE isolates were exposed to systemic daptomycin as it was partially excreted into the gut, highlighting the capacity of the GI tract to serve as a reservoir for the development of antibiotic resistance, even at low antibiotic concentrations (94). Point mutations in the cardiolipin synthase-encoding gene cls were detected in four out of five of these isolate pairs.

Antibiotic resistance: Genetics

In some enterococcal strains, such as vancomycin-resistant E. faecalis V583, acquired genetic elements comprise 25% of the genome (95). There are two major types of plasmids in Enterococci: pheromone-responsive and transposon-type. The pheromone-responsive plasmid pMG2200 encodes VanB-type vancomycin resistance (96). VanA-encoding pheromone-responsive plasmids can be transferred between E. faecium and E. faecalis (97). Large regions of the E. faecalis genome can be shuttled between isolates in vitro via conjugative plasmids, involving up to a quarter of the chromosome (98). Crossover between chromosomal and plasmid DNA can occur through insertion sequences (also known as IS elements). In E. faecalis, pheromone-responsive conjugative plasmids that contain IS256 copies can integrate into the chromosome of recipient strains in vitro and transfer chromosomal DNA from donor isolates, creating hybrid genomes (98). This plasticity of the enterococcal genome has important clinical implications. For example, the transfer of DNA among Enterococci has led to multiple lineages of mutated pbp5 genes conferring ampicillin resistance in hospital-associated strains (84, 85).

Transposons occur throughout the enterococcal genome and are of three types: conjugative, Tn3-family, and composite (flanking IS sequences). The vanA gene cluster is encoded by a Tn3-derivative transposon Tn1546 (99). Tn916-family conjugative transposons include Tn5382 and Tn1549, which are the main genetic elements that contain the VanB resistance operon (100, 101, 102). The gene encoding PBP5 can also be transferred between enterococcal isolates with the Tn916-family conjugative transposon Tn5386 that carries the VanB cluster (103).

DIVERSITY IN GENOMIC COMPOSITION AND HABITATS

The genomic diversity seen among enterococcal strains has been well-characterized by application of high-throughput whole genome sequencing. The first enterococcal genome published in 2002 belonged to E. faecalis V583 (95). Now, there are hundreds of completed or draft genomes available (104). The GC content of enterococcal species can vary from 37% to 45%, and genome sizes can range from 2.7 Mb to 3.6 Mb (105, 106, 107). Compared to commensal enterococcal strains, multidrug-resistant clinical isolates possess larger genomes, through the acquisition of foreign genetic material (107). Hospital-associated E. faecalis strains generally lack CRISPR-Cas systems that help block phage infections and cleave plasmid-encoded DNA (107, 108). In 48 E. faecalis strains, the absence of a CRISPR-Cas system was significantly correlated with resistance to two or more antibiotics (108). Multi-drug resistant E. faecium isolates are also generally CRISPR-Cas deficient, although this relationship has been demonstrated in smaller studies (108, 109). IS elements, such as IS16, drive genomic variation across isolates and likely aided hospital adaptation of Enterococci as they transitioned from antibiotic sensitive to resistant (110, 111, 112). Additionally, recombination has been an important mechanism for generating diversity (89, 98, 113). By contrast, commensals are far less diverse; for example, E. faecalis OG1RF does not contain any laterally-acquired mobile elements and harbors a CRISPR locus (114).

Population genetics

Phylogenetic analyses have found considerable genomic differences between human commensal Enterococci and endemic hospital strains. Nosocomial strains are in fact more closely related to animal isolates than human commensals (107, 115, 116, 117). Whole genome sequencing of E. faecium isolates has revealed two major clades, one comprised of community-derived isolates from healthy humans (clade B) and the other a complex cluster of animal-derived as well as hospital-associated strains (clade A). This split between clades occurred an estimated 3,000 years ago, which coincides roughly with the development of agriculture and animal domestication that conceivably separated animal and human commensals into distinct lineages (117). A second bifurcation occurred almost 75 years ago within clade A between modern nosocomial strains and animal-derived isolates (117). Ampicillin-resistant strains are seen more frequently in pets than in healthy humans (118). Enterococcal strains of animal origin can act as a reservoir of antibiotic resistance elements that can be shared with human isolates (119, 120). For example, VanA genes from animal-derived Enterococci can be laterally transferred to human commensals in the gut (121, 122).

What is the evolutionary relationship between clinical enterococcal isolates? Numerous studies have employed multilocus sequence typing (MLST) as a technique to resolve the enterococcal population structure (123). The process relies on sequencing amplified fragments of seven housekeeping genes (113, 124). Initial studies of E. faecium based on MLST found a distinct cluster of isolates that were enriched in hospitalized patients, named clonal complex 17 (CC17) (89). E. faecalis isolates derived from the hospital environment also group together by MSLT, namely into clonal complexes C2 and C9, which possess more resistance elements and pathogenicity island genes than other clusters (113, 125, 126, 127). However, clinical E. faecium isolates grouped in CC17 are not strictly clonal (111). In phylogenetic analyses that rely on the algorithm eBURST, spurious groupings can occur for species with high recombination rates like E. faecium (128). Analyses of E. faecium strains employing Bayesian models found three major hospital-associated lineages, indicating that nosocomial isolates do not stem from a single ancestral strain (116). Rather, adaptive traits that characterize clinical isolates were likely acquired independently in different genetic backgrounds. Evidence that hospital-associated isolates derived from multiple lineages can also be seen by analyzing the sequence of a single resistance element. Specific amino acid changes in the PBP5 protein are shared between isolates from different sequence types (STs), and sequence variation was found within STs (85). These data indicate that antibiotic resistance developed on the background of multiple enterococcal strains that were poised for survival in the hospital setting.

Habitats

As previously stated, the GI tract is the primary habitat for Enterococci. In animals, E. faecalis, E. faecium, E. hirae, and E. durans are the enterococcal species found most commonly in the gut microbiota (129). Comparisons of VRE in animals and humans have found strains to be host-specific (130). However, patient isolates have been detected in animals such as dogs and pigs, and as discussed above, hospital-adapted strains share a relatively recent close evolutionary relationship to animal isolates (76). While the GI tract represents the largest reservoir for Enterococci, strains have also been found in the environment. It is thought that soil and water isolates are derived from fecal contamination (6, 131, 132, 133). Enterococci possess the ability to adapt to extraintestinal environments, as discussed with regard to the hospital. E. faecalis can survive in nutrient-poor environments, such as sterilized waste for up to 12 days (134). Enterococci are frequently found in human sewage, particularly outside hospitals (135). Not surprisingly, enterococcal strains isolated from effluents are antibiotic resistant. Isolates cultured from sewage as early as the 1970s that were resistant to tetracycline (136). In water, Enterococci are used by the EPA, in addition to total coliform bacteria, as a marker of fecal contamination, after finding a correlation between swimmers’ risk of GI infection and the number of Enterococci cultured from the water site (137). In 2012, 24% of bodies of surface water were classified as impaired in the United States, a number of them due to Enterococci (133).

In the human GI tract, Enterococci live in the small and large intestine. Enterococcal strains represent roughly 1% of human fecal flora, with E. faecalis and E. faecium as the most common inhabitants (15, 16, 17, 18, 19). Average Enterococci density in the GI tract is between 10^4 and 10^6 bacteria per gram wet weight, with E. faecalis found at a somewhat higher abundance than E. faecium (138, 139). However, in one study, E. faecalis was found in over 75% of fecal samples, while E. faecium was detected in 100% (140).

Intestinal commensals thrive in a finely tuned microbial ecology that has evolved over millenia, aiding in nutrient breakdown and the development of mucosal immunity (141, 142). Early-colonizing strains of commensal Enterococci have been shown to contribute to colonic homeostasis through PPARγ1-induced IL-10 and TGF-B expression in vitro and can reduce the severity of infectious diarrhea in children (143, 144, 145). Perturbations to the intestinal microbiota disrupt this symbiotic relationship established with our microbial inhabitants, with important health consequences. Susceptibility to infections is the most well-documented pathology to result from changes in the microbiota, particularly in the context of antibiotic treatment, as detailed in the following section.

INTERACTIONS WITH THE INTESTINAL MICROBIOME

Colonization resistance mediated by the intestinal microbiota

The intestinal microbiota of healthy individuals is comprised of a diverse consortium of bacteria (17, 146, 147). Individuals harbor a range of bacterial compositions, consisting of hundreds of different microbial strains in the colon that mainly fall into the two major phyla, gram-negative Bacteroidetes and gram-positive Firmicutes (17, 148, 149). In addition to variations among individuals, differences in community structure are also found across body sites that exhibit different levels of stability over time, such as between the stable lower (fecal) and variable upper (oral) regions of alimentary canal (150).

As previously noted, administration of broad-spectrum antibiotics allows drug-resistant strains such as VRE to expand dramatically in the gut by perturbing this sensitive microbial ecosystem (151, 22, 23, 24). VRE can expand to 99% of the intestinal lumen’s microbiota in both antibiotic-treated mice and hospitalized patients (23). This overwhelming colonization is associated with translocation into the bloodstream and resulting VRE bacteremia (23, 24). In fact in allo-HSCT patients, VRE colonization was found in over one-third of recipients, and these dominated patients had a 9-fold greater risk for VRE bacteremia (24). This risk persists over time; ampicillin administration leaves mice susceptible to VRE colonization for up to four weeks post-treatment and VRE stably persists in the cecum for at least 60 days (23). In patients, resistant Enterococci can persist for years after antibiotic exposure (152).

The concept of colonization resistance refers to the microbiota’s ability to prevent the entry and growth of exogenous bacteria within its established, complex community (15, 153). Antibiotic treatment abrogates colonization resistance by depleting large swaths of intestinal commensal microorganisms, particularly anaerobic bacteria, that mediate this defense (15, 154, 155, 156). Obligate anaerobes, such as members of the Barnsiella genus and Clostridium cluster XIVa, are highly correlated with intestinal VRE clearance following fecal microbial transplantation (156, 157). How obligate anaerobes provide a robust defense against invading VRE has not been fully elucidated. However, there are broad mechanisms that commensals can employ to exert colonization resistance and prevent infection: 1) indirect elimination that relies on stimulating innate mucosal immunity, 2) continual maintenance of mucosal barrier integrity, and 3) direct antagonism.

Indirect inhibition through innate immune defense

Intestinal microbes can stimulate innate receptors on immune cells and induce the production of antimicrobial peptides (AMPs) in other intestinal cell types. Paneth cells and intestinal epithelial cells produce RegIIIγ, a C-type lectin driven by TLR signaling with bactericidal activity against gram-positive bacteria (158, 159, 160). Secreted RegIIIγ kills bacteria by binding to peptidoglycans of the bacterial cell wall and forming pores (161). Antibiotic treatment reduces expression of RegIIIγ and, in mice, increases susceptibility to VRE colonization and bacteremia (162). Oral administration of LPS mimics commensal microbial signals and restores RegIIIγ production, thereby increasing resistance to VRE (162). A signaling pathway driving RegIIIγ expression was delineated by administration of the bacterial TLR5 ligand, flagellin. Flagellin administered intravenously stimulates the CD103+ CD11b+ subset of dendritic cells to produce IL-23, which drives the IL-22-mediated production of RegIIIγ by intestinal epithelial cells (163). Commensals can thus work in concert with the mucosal immune system to suppress VRE outgrowth within the intestinal ecosystem.

Indirect inhibition through intestinal barrier maintenance

Intestinal microbes are separated from the mucosal epithelium and its distal lamina propria by mucus that coats the epithelial surface. The colonic epithelium is covered by a dense 50-um thick inner mucin layer composed primarily of Muc2 and a less dense outer stratum (164). Maintenance of a healthy epithelial barrier and intact gut physiology, such as gastric acid production, inhibits bacterial colonization of the GI tract (165). Goblet cells produce mucin, and secretion is stimulated by commensal bacteria in a MyD88-dependent manner (166, 167, 168). Following antibiotic treatment, the mucin layer thins; without a robust physical barrier, intestinal microbes can directly access and potentially breach the epithelium (169). Both the density and composition of the mucus layers limits bacterial invasion. RegIIIγ is associated with mucin and reduces the density of intestinal bacteria near epithelial cells (170, 171, 172).

Compared to other antibiotic-resistant pathogens such as Klebsiella pneumoniae (KP), VRE is spatially segregated from the intestinal mucus layer and adjacent epithelium even after antibiotic treatment with its notable mucin reduction (173). Visualization of the colonic lumen reveals that VRE does not infiltrate the inner mucin layer and, despite high luminal density, very few bacteria translocate to the mesenteric lymph nodes (MLN) (173). Interestingly, co-colonization of mice with VRE and KP, which can more deeply penetrate the mucus coating, enables VRE to gain access to the MLNs, possibly by KP-induced alterations to the mucin composition (173). Intact mucin production, which is in part regulated by commensal microbes, likely limits the invasive potential of intestinal Enterococci.

Direct inhibition by anaerobic commensals

In the first study of its kind for VRE, a defined consortium of commensals was identified as capable of restoring colonization resistance in mice (157). Antibiotic-treated mice were orally administrated diluted doses of fecal microbiota from a colony of mice that had received ampicillin for over fifteen years. Bacterial isolates in low-dose fractions that conferred resistance to VRE were identified, cultured, and administrated in discrete combinations to mice maintained on ampicillin. Through a series of leave-one-out adoptive transfers, a minimum of four anaerobic isolates were found to successfully prevent and clear VRE from the gut: Blautia producta, Clostridium bolteae, Bacteroides sartorii, and Parabacteroides distasonis (157). Of the four-commensal mixture, Blautia producta was shown ex vivo as the member that directly inhibit VRE growth, although the exact mechanism remains unknown. One possible mechanism of inhibition is through the production of toxic substances such as bacteriocins, which are small molecules with antimicrobial activity. Lactococcus lactis strains engineered to express bacteriocins significantly inhibited VRE growth in vitro (174). Oral administration of bacteriocin-producing Lactococcus lactis MM19 eliminated VRE at a faster rate from the gut of mice than mock treatment (175).

Direct inhibition by commensal Enterococci

Recent studies have examined the colonization dynamics between enterococcal commensals and nosocomial isolates in the GI tract. While resistant isolates outcompete sensitive Enterococci in the context of antibiotic pressure, intestinal colonization in patients declines following discharge (176). In in vivo competition assays that compared the colonization ability of E. faecium strains in antibiotic-treated mice, isolates from clade B (commensal-associated) outcompeted those from subclade A1 (hospital-derived) after two weeks (177).

Commensal Enterococci have developed sophisticated defense mechanisms to eliminate exogenous enterococcal competitors from the gut. Bacteriocin-coding genes are commonly harbored on plasmids in Enterococci. Commensal E. faecalis that express a pheromone-responsive conjugative plasmid encoding bacteriocin bac-21 outcompeted VRE lacking it (178). This plasmid, pPD1, is also quickly transferred to naive intestinal commensals by conjugation (178). Pheromones are secreted short lipoprotein signal peptide fragments that act as chemical messengers between bacteria and can mediate cell death. The multi-drug resistant E. faecalis isolate V583 harbors a plasmid pTEF2 that renders it susceptible to a killing mechanism induced by commensal-derived pheromone cOB1 (179). Bacteriophages are viruses that selectively infect and kill microbes. Given their selective killing, phages could be used therapeutically as a narrow-spectrum antimicrobial. E. faecalis strains that contain the bacteriophage ϕV1/7 in their genetic repertoire possess a growth advantage over related bacteria that lack it through phage-mediated lysis of competitors (180). In a mouse model of VRE bacteremia, intraperintoneal injection of ENB6 phage protected all mice when administered shortly after lethal VRE challenge and half of the mice when administered after the mice were moribund (181).

Enterococci as probiotics

The benefits of using Enterococci as probiotics have been controversial (182). Given the capacity of enterococcal isolates to share mobile virulence elements in the gut, there is concern of spreading antibiotic resistance if carried or obtained by probiotics. However, enterococcal strains such as E. faecium SF68 and E. faecalis Symbio-flor have been marketed as probiotics for two decades without incidence and with very few reported adverse events (182, 183, 184). Enterococcal probiotics have been shown to be effective in limiting gastrointestinal infectious burden. A Cochrane meta-review of the literature found E. faecium SF68 to be an efficacious treatment of GI infections (184). Inoculation of the E. faecium SF68 alone to adults and children with enteritis reduces the length of illness (182, 184, 185, 186). A probiotic mix containing E. faecalis as well as Bacillus mesentericus and Clostridium butyricum shortened the severity and duration of infectious diarrhea in children (145). In studies on diarrhea lasting 4 days or more, live Lactobacillus casei strain GG had a larger treatment effect size (0.59) than live Enterococcus SF68 (0.2), although the former had nearly twice as many participants enrolled in all trials (184).

Fecal microbiota transplantation and probiotics as treatment for VRE colonization

Given the rise of antibiotic resistance, fecal microbiota transplantation (FMT) is an attractive alternative therapy to treat antibiotic-resistant pathogens and an area of active research. FMT is remarkably successful at curing chronic, intractable C. difficile infection (187). A secondary analysis of a study involving patients with recurrent C. difficile infection showed that a human-derived FMT can reduce VRE colonization (188). However, the risk of unwittingly transmitting pathogenic microorganisms through FMTs is not insignificant, especially since many constituents of the microbiota have only recently been identified, if not characterized. This concern is particularly relevant to patients colonized with VRE, who are often immunocompromised. The field is actively exploring methods to perfect the acquisition of transferred bacteria and define critical members of FMTs that target infectious agents (189, 190).

To date, clinical trials on the impact of probiotics on the intestinal VRE carriage are limited. In a randomized study of 21 renal patients harboring VRE in their GI tract, ingestion of a yogurt supplemented with Lactobacillus rhamnosus GG reduced VRE density to the limit of detection in all patients receiving the probiotic (191). VRE burden decreased during a three-week oral supplementation with L. rhamnosus GG in a randomized clinical trial of 61 children (192). This effect was not seen with five-week administration of L. rhamnosus Lcr35 in a randomized study of nine patients (193). A two-week course of L. rhamnosus GG administration in 11 patients with comorbidities also did not affect VRE colonization (194). Studies of enterococcal probiotics have failed to demonstrate their potential to limit drug-resistant Enterococci colonization. In a prospective cohort study with over 500 hospitalized patients, a 10-strain mixture that contained E. faecium and numerous Lactobacillus isolates did not prevent ampicillin-resistant E. faecium acquisition (195).

The optimal design of probiotic consortia utilizes preclinical mouse models for candidate screening and follow-up mechanistic studies. Microbiome research relies on deep 16S rRNA gene and shotgun sequencing to profile bacterial communities of the gut and to predict candidate commensals that confer colonization resistance in time-series microbiota-reconstitution experiments. Ecological modeling of the microbiota using 16S sequencing data accurately predicted fluctuations in the composition of the microbiota following clindamycin administration and C. difficile colonization, and proposed the anaerobe Coprobacillus as a commensal capable of inhibiting Enterococcal growth (196). In vivo adoptive transfer experiments allow investigators to further elucidate the mechanisms of colonization resistance provided by reconstituted commensals. In a mouse model of C. difficile infection, Clostridium scindens protected antibiotic-treated mice from C. difficile colonization in by restoring secondary bile salt levels that inhibit the pathogen’s growth (190). How these findings are best translated to treating at-risk patients is yet to be determined. In a promising phase 1b trial, orally administered capsules of 50 human-derived live Firmicutes spores prevented recurrent C. difficile infection, while the phase II clinical study found no efficacy (197, 198). A key question facing the translation of optimal bacterial combinations into patient therapy is what is required for a high transplantation efficacy. The study that defined a minimal consortium for VRE in mice highlights this challenge (157). Successful colonization of Blautia producta in ampicillin-treated mice required the adoptive transfer of three additional commensals. Bacteroides sartorii and Parabacteroides distasonis inactivate ampicillin through the production of β-lactamase, which was critical for ampicillin-sensitive isolates’ survival in the GI tract, while Clostridium bolteae supported Blautia producta’s engraftment through an unknown mechanism (157). Modulating the local gut environment through drug inactivation with probiotics is of particular importance for preventing VRE colonization in patients currently receiving antibiotics (199). Probiotic commensals can limit pathogen colonization in the gut by mitigating the disruptive effects of antibiotics to begin with. A Bacteroides thetaiotaomicron strain that produces a cephalosporinase has been shown to prevent intestinal VRE outgrowth by inactivating ceftriaxone and thus mitigating any significant changes to the microbiota (200).

Another open question is whether a protective microbial consortium should be tailored to individual patients, and if so, how to scale such a design. Given the falling costs of deep sequencing, profiling patients’ microbiota may occur regularly in clinical practice. In the context of VRE, patients with different degrees of immune system impairment and treatment histories may benefit from personalized alterations to the minimally-defined protective consortium. For example, patients who recently received antibiotics may be deficient in nutrients that resistance-mediating bacteria require to survive in the gut, necessitating additional isolates to support successful engraftment. Mouse models would not be a scalable approach to test these individual modifications. In this era of deep sequencing, we can potentially integrate diet, treatment regimens and gut microbiome data to build machine-learning algorithms that can assess a patient’s risk of VRE colonization and optimize probiotic combinations. Incorporating information on microbiome composition and function improved predictions for individuals’ glycemic response following a meal and helped design dietary interventions for better glycemic control (201). Such data-driven approaches may help tailor preclinical findings to individual patients at scale to successfully mitigate their susceptibility to VRE colonization.

References

  • 1.Euzeby JP. List of bacterial names with standing in nomenclature: a folder available on the internet. Parte AC, editor. Int. J. Syst. Bacteriol. 1997;47:590–592. doi: 10.1099/00207713-47-2-590. Retrieved August 22 2016: http://www.bacterio.net/enterococcus.html. [DOI] [PubMed]
  • 2.Mundt JO. Occurrence of enterococci in animals in a wild environment. Appl Microbiol. 1963;11:136–40. doi: 10.1128/am.11.2.136-140.1963. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.MacCallum WG, Hastings TW. A case of acute endocarditis caused by Micrococcus zymogenes (Nov. Spec.), with a description of the microorganism. The Journal of Experimental Medicine. 1899;4(56):521–534. doi: 10.1084/jem.4.5-6.521. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Thiercelin ME, Jouhaud L. Sur un diplococque saprophyte de l'intestin susceptible de devenir pathogene. CR Soc. Biol. 1899;5:269–271. [Google Scholar]
  • 5.Schleifer KH, Kilpper-Bälz R. Transfer of Streptococcus faecalis and Streptococcus faecium to the genus Enterococcus nom. Rev. as Enterococcus faecalis comb. Nov. and Enterococcus faecium comb. Nov. International Journal of Systematic and Evolutionary Microbiology. 1984;34(1):31–34. [Google Scholar]
  • 6.Sherman JM. The Streptococci. Bacteriological Reviews. 1937;1(1):3. doi: 10.1128/br.1.1.3-97.1937. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Facklam RR. Comparison of several laboratory media for presumptive identification of Enterococci and group D Streptococci. Appl Microbiol. 1973;26(2):138–45. doi: 10.1128/am.26.2.138-145.1973. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Facklam RR, Collins MD. Identification of enterococcus species isolated from human infections by a conventional test scheme. J Clin Microbiol. 1989;27(4):731–4. doi: 10.1128/jcm.27.4.731-734.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Facklam RR, Carvalho MG, Teixeira LM. Enterococcus. In: Gilmore MS, Clewell DB, Courvalin P, Dunny GM, Murray BE, Rice LB, editors. The Enterococci: Pathogenesis, Molecular Biology, and Antibiotic Resistance. Washington, DC: ASM Press; 2002. [Google Scholar]
  • 10.Martin JD, Mundt JO. Enterococci in insects. Appl Microbiol. 1972;24(4):575–580. doi: 10.1128/am.24.4.575-580.1972. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Thiercelin ME. Morphologie et modes de reproduction de l'enterocoque. Comptes Rendus des Seances de la Societe de Biologie et des ses Filiales. 1899;11:551–553. [Google Scholar]
  • 12.Andrewes FW, Horder TJ. A study of Streptococci pathogenic for man. The Lancet. 1906;168(4335):852–855. [Google Scholar]
  • 13.Orla-Jensen S. The lactic acid bacteria. Memoirs of the Academy of the Royal Society of Denmark. Section of Sciences Series. 1919;85:81–197. [Google Scholar]
  • 14.Gilmore MS, Lebreton F, van Schaik W. Genomic transition of Enterococci from gut commensals to leading causes of multidrug-resistant hospital infection in the antibiotic era. Curr Opin Microbiol. 2013;16(1):10–6. doi: 10.1016/j.mib.2013.01.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Vollaard EJ, Clasener HA. Colonization resistance. Antimicrob Agents Chemother. 1994;38(3):409–414. doi: 10.1128/aac.38.3.409. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Sghir A, Gramet G, Suau A, Rochet V, Pochart P, Dore J. Quantification of bacterial groups within human fecal flora by oligonucleotide probe hybridization. Appl Environ Microbiol. 2000;66(5):2263–6. doi: 10.1128/aem.66.5.2263-2266.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Eckburg PB, et al. Diversity of the human intestinal microbial flora. Science. 2005;308(5728):1635–8. doi: 10.1126/science.1110591. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Tendolkar PM, Baghdayan AS, Shankar N. Pathogenic Enterococci: New developments in the 21st century. Cell Mol Life Sci. 2003;60(12):2622–36. doi: 10.1007/s00018-003-3138-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Lebreton F, Willems RJL, Gilmore MS. Enterococcus Diversity, Origins in Nature, and Gut Colonization. In: Gilmore MS, Clewell DB, Ike Y, et al., editors. Enterococci: From Commensals to Leading Causes of Drug Resistant Infection [Internet] Boston: 2014. [Massachusetts Eye and Ear Infirmary]. 2014-. Available from: http://www.ncbi.nlm.nih.gov/books/NBK190427/ [PubMed] [Google Scholar]
  • 20.Huycke MM, Sahm DF, Gilmore MS. Multiple-drug resistant Enterococci: The nature of the problem and an agenda for the future. Emerg Infect Dis. 1998;4(2):239–49. doi: 10.3201/eid0402.980211. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Galloway-Peña JR, Nallapareddy SR, Arias CA, Eliopoulos GM, Murray BM. Analysis of clonality and antibiotic resistance among early clinical isolates of Enterococcus faecium in the United States. J Infect Dis. 2009;200(10):1566–73. doi: 10.1086/644790. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Donskey CJ, et al. Effect of antibiotic therapy on the density of vancomycin-resistant Enterococci in the stool of colonized patients. N Engl J Med. 2000;343(26):1925–32. doi: 10.1056/NEJM200012283432604. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Ubeda C, et al. Vancomycin-resistant Enterococcus domination of intestinal microbiota is enabled by antibiotic treatment in mice and precedes bloodstream invasion in humans. J Clin Invest. 2010;120(12):4332–41. doi: 10.1172/JCI43918. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Taur Y, et al. Intestinal domination and the risk of bacteremia in patients undergoing allogeneic hematopoietic stem cell transplantation. Clin Infect Dis. 2012;55(7):905–14. doi: 10.1093/cid/cis580. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Hidron AI, Edwards JR, Patel J, Horan TC, Sievert DM, Pollock DA, Fridkin SC. Antimicrobial-resistant pathogens associated with healthcare-associated infections: Annual summary of data reported to the National Healthcare Safety Network at the Centers for Disease Control and Prevention, 2006--2007. Infect Control Hosp Epidemiol. 2008;29(11):996–1011. doi: 10.1086/591861. [DOI] [PubMed] [Google Scholar]
  • 26.Jett BD, Huycke MM, Gilmore MS. Virulence of Enterococci. Clin Microbiol Rev. 1994;7(4):462–478. doi: 10.1128/cmr.7.4.462. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.CDC (Centers for Disease Control and Prevention) Antibiotic resistance threats in the United States. 2013 http://www.cdc.gov/drugresistance/threat-report-2013/
  • 28.Agudelo Higuita NI, Huycke MM. Enterococcal Disease, Epidemiology, and Implications for Treatment. In: Gilmore MS, Clewell DB, Ike Y, et al., editors. Enterococci: From Commensals to Leading Causes of Drug Resistant Infection [Internet] Boston: Massachusetts Eye and Ear Infirmary; 2014. 2014-. Available from: https://www.ncbi.nlm.nih.gov/books/NBK190429/ [PubMed] [Google Scholar]
  • 29.Weiner LM, Amy K, Webb AK, Limbago B, Dudeck MA, Patel J, Kallen AJ, Edwards JR, Sievert DM. Antimicrobial-resistant pathogens associated with healthcare-associated infections: Summary of data reported to the National Healthcare Safety Network at the Centers for Disease Control and Prevention, 2011–2014. Infect Control Hosp Epidemiol. 2016:1–14. doi: 10.1017/ice.2016.174. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Wisplinghoff H, Bischoff T, Tallent SM, Seifert H, Wenzel RP, Edmond MB. Nosocomial bloodstream infections in US hospitals: Analysis of 24,179 cases from a prospective nationwide surveillance study. Clin Infect Dis. 2004;39(3):309–17. doi: 10.1086/421946. [DOI] [PubMed] [Google Scholar]
  • 31.Berg RD. Bacterial translocation from the gastrointestinal tract. Adv Exp Med Biol. 1999;473:11–30. doi: 10.1007/978-1-4615-4143-1_2. [DOI] [PubMed] [Google Scholar]
  • 32.Kuehnert MJ, Jernigan JA, Pullen AL, Rimland D, Jarvis WR. Association between mucositis severity and vancomycin-resistant Enterococcal bloodstream infection in hospitalized cancer patients. Infect Control Hosp Epidemiol. 1999;20(10):660–3. doi: 10.1086/501561. [DOI] [PubMed] [Google Scholar]
  • 33.Roghmann MC, McCarter RJ, Brewrink J, Cross AS, Morris JG. Clostridium difficile infection is a risk factor for bacteremia due to vancomycin-resistant Enterococci (VRE) in VRE-colonized patients with acute leukemia. Clin Infect Dis. 1997;25(5):1056–9. doi: 10.1086/516112. [DOI] [PubMed] [Google Scholar]
  • 34.Lautenbach E, Bilker WB, Brennan PJ. Enterococcal bacteremia: Risk factors for vancomycin resistance and predictors of mortality. Infect Control Hosp Epidemiol. 1999;20(5):318–23. doi: 10.1086/501624. [DOI] [PubMed] [Google Scholar]
  • 35.Murdoch DR, et al. Clinical presentation, etiology, and outcome of infective endocarditis in the 21st century: The international collaboration on endocarditis-prospective cohort study. Arch Intern Med. 2009;169(5):463–73. doi: 10.1001/archinternmed.2008.603. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Anderson DJ, Murdoch DR, Sexton DJ, Reller LB, Stout JE, Cabell CH, Corey GR. Risk factors for infective endocarditis in patients with Enterococcal bacteremia: A case-control study. Infection. 2004;32(2):72–7. doi: 10.1007/s15010-004-2036-1. [DOI] [PubMed] [Google Scholar]
  • 37.Arias CA, Murray BE. The rise of the Enterococcus: Beyond vancomycin resistance. Nat Rev Microbiol. 2012;10(4):266–78. doi: 10.1038/nrmicro2761. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Mandell GL, Kaye D, Levison ME, Hook EW. Enterococcal endocarditis. An analysis of 38 patients observed at the New York Hospital-Cornell Medical Center. Arch Intern Med. 1970;125(2):258–64. doi: 10.1001/archinte.125.2.258. [DOI] [PubMed] [Google Scholar]
  • 39.D'Agata EM, Green WK, Schulman G, Li H, Tang YW, Schaffner W. Vancomycin-resistant Enterococci among chronic hemodialysis patients: A prospective study of acquisition. Clin Infect Dis. 2001;32(1):23–9. doi: 10.1086/317549. [DOI] [PubMed] [Google Scholar]
  • 40.Boyce JM, et al. Outbreak of multidrug-resistant Enterococcus faecium with transferable vanB class vancomycin resistance. J Clin Microbiol. 1994;32(5):1148–53. doi: 10.1128/jcm.32.5.1148-1153.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Boyce JM, Mermel LA, Zervos MJ, Rice LB, Potter-Bynoe G, Giorgio C, Medeiros AA. Controlling vancomycin-resistant Enterococci. Infect Control Hosp Epidemiol. 1995;16(11):634–637. doi: 10.1086/647028. [DOI] [PubMed] [Google Scholar]
  • 42.Handwerger S, Raucher B, Altarac D, Monka J, Marchione S, Singh KV, Murray BE, Wolff J, Walters B. Nosocomial outbreak due to Enterococcus faecium highly resistant to vancomycin, penicillin, and gentamicin. Clin Infect Dis. 1993;16(6):750–5. doi: 10.1093/clind/16.6.750. [DOI] [PubMed] [Google Scholar]
  • 43.Cetinkaya Y, Falk P, Mayhall CG. Vancomycin-resistant Enterococci. Clin Microbiol Rev. 2000;13(4):686–707. doi: 10.1128/cmr.13.4.686-707.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Howden BP, et al. Genomic insights to control the emergence of vancomycin-resistant Enterococci. MBio. 2013;4(4) doi: 10.1128/mBio.00412-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Hota B. Contamination, disinfection, and cross-colonization: Are hospital surfaces reservoirs for nosocomial infection? Clin Infect Dis. 2004;39(8):1182–9. doi: 10.1086/424667. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Drees M, et al. Prior environmental contamination increases the risk of acquisition of vancomycin-resistant Enterococci. Clin Infect Dis. 2008;46(5):678–85. doi: 10.1086/527394. [DOI] [PubMed] [Google Scholar]
  • 47.Austin DJ, Bonten MJ, Weinstein RA, Slaughter S, Anderson RM. Vancomycin-resistant Enterococci in intensive-care hospital settings: Transmission dynamics, persistence, and the impact of infection control programs. Proceedings of the National Academy of Sciences. 1999;96(12):6908–6913. doi: 10.1073/pnas.96.12.6908. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Muto CA, Jernigan JA, Ostrowsky BE, Richet HM, Jarvis WR, Boyce JM, Farr BM SHEA. SHEA guideline for preventing nosocomial transmission of multidrug-resistant strains of Staphylococcus aureus and Enterococcus. Infect Control Hosp Epidemiol. 2003;24(5):362–86. doi: 10.1086/502213. [DOI] [PubMed] [Google Scholar]
  • 49.Duckro AN, Blom DW, Lyle EA, Weinstein RA, Hayden MK. Transfer of vancomycin-resistant Enterococci via health care worker hands. Arch Intern Med. 2005;165(3):302–7. doi: 10.1001/archinte.165.3.302. [DOI] [PubMed] [Google Scholar]
  • 50.Mayer RA, Geha RC, Helfand MS, Hoyen CK, Salata RA, Donskey CJ. Role of fecal incontinence in contamination of the environment with vancomycin-resistant Enterococci. American Journal of Infection Control. 2003;31(4):221–225. doi: 10.1067/mic.2003.45. [DOI] [PubMed] [Google Scholar]
  • 51.Zervos MJ, Terpenning MS, Schaberg DR, Therasse PM, Medendorp SV, Kauffman CA. High-level aminoglycoside-resistant Enterococci: Colonization of nursing home and acute care hospital patients. Arch Intern Med. 1987;147(9):1591–1594. doi: 10.1001/archinte.147.9.1591. [DOI] [PubMed] [Google Scholar]
  • 52.Bonten MJ, Slaughter S, Ambergen AW, Hayden MK, van Voorhis J, Nathan C, Weinstein RA. The role of "colonization pressure" in the spread of vancomycin-resistant Enterococci: An important infection control variable. Arch Intern Med. 1998;158(10):1127–32. doi: 10.1001/archinte.158.10.1127. [DOI] [PubMed] [Google Scholar]
  • 53.Tornieporth NG, Roberts RB, John J, Hafner A, Riley LW. Risk factors associated with vancomycin-resistant Enterococcus faecium infection or colonization in 145 matched case patients and control patients. Clinical Infectious Diseases. 1996;23(4):767–772. doi: 10.1093/clinids/23.4.767. [DOI] [PubMed] [Google Scholar]
  • 54.Vergis EN, Hayden MK, Chow JW, Snydman DR, Zervos MJ, Linden PK, Wagener MM, Schmitt B, Muder RR. Determinants of vancomycin resistance and mortality rates in Enterococcal bacteremia. A prospective multicenter study. Ann Intern Med. 2001;135(7):484–92. doi: 10.7326/0003-4819-135-7-200110020-00007. [DOI] [PubMed] [Google Scholar]
  • 55.Jawetz E, Sonne M. Penicillin-Streptomycin treatment of Enterococcal endocarditis - a re-evaluation. N Engl J Med 1966. 1966;274:710–715. doi: 10.1056/NEJM196603312741304. [DOI] [PubMed] [Google Scholar]
  • 56.Moellering RC, Wennersten C, Weinberg AN. Studies on antibiotic synergism against Enterococci. I. Bacteriologic studies. J Lab Clin Med. 1971;77(5):821–8. [PubMed] [Google Scholar]
  • 57.Murray BE. The life and times of the enterococcus. Clin Microbiol Rev. 1990;3(1):46–6. doi: 10.1128/cmr.3.1.46. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Nigo M, Munita JM, Arias CA, Murray BE. What's new in the treatment of Enterococcal endocarditis? Curr Infect Dis Rep. 2014;16(10):431. doi: 10.1007/s11908-014-0431-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Munita JM, Arias CA, Murray BE. Editorial Commentary: Enterococcus faecalis infective endocarditis: is it time to abandon aminoglycosides? Clin Infect Dis. 2013;56(9):1269–72. doi: 10.1093/cid/cit050. [DOI] [PubMed] [Google Scholar]
  • 60.Mainardi JL, Gutmann L, Acar JF, Goldstein FW. Synergistic effect of amoxicillin and cefotaxime against Enterococcus faecalis. Antimicrob Agents Chemother. 1995;39(9):1984–7. doi: 10.1128/aac.39.9.1984. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Baddour LM, et al. Infective endocarditis in adults: Diagnosis, antimicrobial therapy, and management of complications: A scientific statement for healthcare professionals from the American Heart Association. Circulation. 2015;132(15):1435–86. doi: 10.1161/CIR.0000000000000296. [DOI] [PubMed] [Google Scholar]
  • 62.O'Driscoll T, Crank CW. Vancomycin-resistant Enterococcal infections: Epidemiology, clinical manifestations, and optimal management. Infect Drug Resist. 2015;8:217–30. doi: 10.2147/IDR.S54125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Treitman AN, Yarnold PR, Warren J, Noskin GA. Emerging incidence of Enterococcus faecium among hospital isolates (1993 to 2002) J Clin Microbiol. 2005;43(1):462–3. doi: 10.1128/JCM.43.1.462-463.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Top J, Willems R, Blok H, de Regt M, Jalink K, Troelstra A, Goorhuis B, Bonten M. Ecological replacement of Enterococcus faecalis by multiresistant clonal complex 17 Enterococcus faecium. Clin Microbiol Infect. 2007;13(3):316–9. doi: 10.1111/j.1469-0691.2006.01631.x. [DOI] [PubMed] [Google Scholar]
  • 65.Kristich CJ, Rice LB, Arias CA. Enterococcal Infection - Treatment and Antibiotic Resistance. In: Gilmore MS, Clewell DB, Ike Y, et al., editors. Enterococci: From Commensals to Leading Causes of Drug Resistant Infection [Internet] Boston: Massachusetts Eye and Ear Infirmary; 2014. 2014-. Available from: http://www.ncbi.nlm.nih.gov/books/NBK190420/ [PubMed] [Google Scholar]
  • 66.Edelsberg J, et al. Prevalence of antibiotic resistance in US hospitals. Diagn Microbiol Infect Dis. 2014;78(3):255–62. doi: 10.1016/j.diagmicrobio.2013.11.011. [DOI] [PubMed] [Google Scholar]
  • 67.Leclercq R, Derlot E, Duval J, Courvalin P. Plasmid-mediated resistance to vancomycin and teicoplanin in Enterococcus faecium. The New England Journal of Medicine. 1988;319(3):157–161. doi: 10.1056/NEJM198807213190307. [DOI] [PubMed] [Google Scholar]
  • 68.Uttley AHC, George RC, Naidoo J, Woodford N, Johnson AP, Collins CH, Morrison D, Gilfillan AJ, Fitch LE, Heptonstall J. High-level vancomycin-resistant Enterococci causing hospital infections. Epidemiology and Infection. 1989;103(01):173–181. doi: 10.1017/s0950268800030478. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Samore MH, Carmeli Y, Eliopoulos GM. Antecedent treatment with different antibiotic agents as a risk factor for vancomycin-resistant enterococcus. Emerg Infect Dis. 2002;8(8):802–7. doi: 10.3201/eid0808.010418. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Currie BP, Lemos-Filho L. Evidence for biliary excretion of vancomycin into stool during intravenous therapy: Potential implications for rectal colonization with vancomycin-resistant Enterococci. Antimicrob Agents Chemother. 2004;48(11):4427–9. doi: 10.1128/AAC.48.11.4427-4429.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Aarestrup FM. Characterization of glycopeptide-resistant Enterococcus faecium (GRE) from broilers and pigs in Denmark: Genetic evidence that persistence of GRE in pig herds is associated with coselection by resistance to macrolides. J Clin Microbiol. 2000;38(7):2774–7. doi: 10.1128/jcm.38.7.2774-2777.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Bates J. Epidemiology of vancomycin-resistant enterococci in the community and the relevance of farm animals to human infection. J Hosp Infect. 1997;37(2):89–101. doi: 10.1016/s0195-6701(97)90179-1. [DOI] [PubMed] [Google Scholar]
  • 73.Van Tyne D, Gilmore MS. Friend turned foe: Evolution of Enterococcal virulence and antibiotic resistance. Annu Rev Microbiol. 2014;68:337–56. doi: 10.1146/annurev-micro-091213-113003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Bager F, Aarestrup FM, Madsen M, Wegener HC. Glycopeptide resistance in Enterococcus faecium from broilers and pigs following discontinued use of avoparcin. Microb Drug Resist. 1999;5(1):53–6. doi: 10.1089/mdr.1999.5.53. [DOI] [PubMed] [Google Scholar]
  • 75.Klare I, Badstübner D, Konstabel C, Böhme G, Claus H, Witte W. Decreased incidence of vanA-type vancomycin-resistant Enterococci isolated from poultry meat and from fecal samples of humans in the community after discontinuation of avoparcin usage in animal husbandry. Microb Drug Resist. 1999;5(1):45–52. doi: 10.1089/mdr.1999.5.45. [DOI] [PubMed] [Google Scholar]
  • 76.Hammerum AM. Enterococci of animal origin and their significance for public health. Clin Microbiol Infect. 2012;18(7):619–25. doi: 10.1111/j.1469-0691.2012.03829.x. [DOI] [PubMed] [Google Scholar]
  • 77.Werner G, et al. Emergence and spread of vancomycin resistance among Enterococci in Europe. Euro Surveill. 2008;13(47) [PubMed] [Google Scholar]
  • 78.Fontana R, Grossato A, Rossi L, Cheng YR, Satta G. Transition from resistance to hypersusceptibility to beta-lactam antibiotics associated with loss of a low-affinity penicillin-binding protein in a Streptococcus faecium mutant highly resistant to penicillin. Antimicrob Agents Chemother. 1985;28(5):678–683. doi: 10.1128/aac.28.5.678. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Williamson R, Le Bouguenec C, Gutmann L, Horaud T. One or two low affinity penicillin-binding proteins may be responsible for the range of susceptibility of Enterococcus faecium to benzylpenicillin. Microbiology. 1985;131(8):1933–1940. doi: 10.1099/00221287-131-8-1933. [DOI] [PubMed] [Google Scholar]
  • 80.Fontana R, Cerini R, Longoni P, Grossato A, Canepari P. Identification of a streptococcal penicillin-binding protein that reacts very slowly with penicillin. J Bacteriol. 1983;155(3):1343–50. doi: 10.1128/jb.155.3.1343-1350.1983. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Rybkine T, Mainardi J, Sougakoff W, Collatz E, Gutmann L. Penicillin-binding protein 5 sequence alterations in clinical isolates of Enterococcus faecium with different levels of β-lactam resistance. Journal of Infectious Diseases. 1998;178(1):159–163. doi: 10.1086/515605. [DOI] [PubMed] [Google Scholar]
  • 82.Zorzi W, Zhou WY, Dardenne O, Lamotte J, Raze J, Pierre J, Gutmann L, Coyette J. Structure of the low-affinity penicillin-binding protein 5 pbp5fm in wild-type and highly penicillin-resistant strains of Enterococcus faecium. J Bacteriol. 1996;178(16):4948–4957. doi: 10.1128/jb.178.16.4948-4957.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Rice LB, Hutton-Thomas R, Viera Lakticova V, Helfand MS, Donskey CJ. Β-lactam antibiotics and gastrointestinal colonization with vancomycin-resistant Enterococci. Journal of Infectious Diseases. 2004;189(6):1113–1118. doi: 10.1086/382086. [DOI] [PubMed] [Google Scholar]
  • 84.Rice LB, Carias LL, Rudin S, Lakticová V, Wood A, Hutton-Thomas R. Enterococcus faecium low-affinity pbp5 is a transferable determinant. Antimicrob Agents Chemother. 2005;49(12):5007–12. doi: 10.1128/AAC.49.12.5007-5012.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Galloway-Peña JR, Rice LB, Murray BE. Analysis of PBP5 of early U.S. isolates of Enterococcus faecium: Sequence variation alone does not explain increasing ampicillin resistance over time. Antimicrob Agents Chemother. 2011;55(7):3272–7. doi: 10.1128/AAC.00099-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Grayson ML, Eliopoulos GM, Wennersten CB, Ruoff KL, De Girolami PC, Ferraro MJ, Moellering RC. Increasing resistance to beta-lactam antibiotics among clinical isolates of Enterococcus faecium: A 22-year review at one institution. Antimicrob Agents Chemother. 1991;35(11):2180–4. doi: 10.1128/aac.35.11.2180. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Arthur M, Courvalin P. Genetics and mechanisms of glycopeptide resistance in Enterococci. Antimicrob Agents Chemother. 1993;37(8):1563–71. doi: 10.1128/aac.37.8.1563. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Courvalin P. Vancomycin resistance in gram-positive cocci. Clinical Infectious Diseases. 2006;42(Supplement 1):S25–S34. doi: 10.1086/491711. [DOI] [PubMed] [Google Scholar]
  • 89.Willems RJ, Top J, Van Santen M, Robinson DA, Coque TM, Baquero F, Grundmann H, Bonten MJ. Global spread of vancomycin-resistant Enterococcus faecium from distinct nosocomial genetic complex. Emerg Infect Dis. 2005;11(6):821–828. doi: 10.3201/eid1106.041204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Alborn WE, Allen NE, Preston DA. Daptomycin disrupts membrane potential in growing Staphylococcus aureus. Antimicrob Agents Chemother. 1991;35(11):2282–7. doi: 10.1128/aac.35.11.2282. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Silverman JA, Perlmutter NG, Shapiro HM. Correlation of daptomycin bactericidal activity and membrane depolarization in Staphylococcus aureus. Antimicrob Agents Chemother. 2003;47(8):2538–2544. doi: 10.1128/AAC.47.8.2538-2544.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Arias CA, et al. Genetic basis for in vivo daptomycin resistance in Enterococci. New England Journal of Medicine. 2011;365(10):892–900. doi: 10.1056/NEJMoa1011138. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Palmer KL, Kelli L, Daniel A, Hardy C, Silverman J, Gilmore MS. Genetic basis for daptomycin resistance in Enterococci. Antimicrob Agents Chemother. 2011;55(7):3345–56. doi: 10.1128/AAC.00207-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Lellek H, Franke GC, Ruckert C, Wolters M, Wolschke C, Christner M, Büttner H, Malik Alawi M, Kröger N, Rohde H. Emergence of daptomycin non-susceptibility in colonizing vancomycin-resistant Enterococcus faecium isolates during daptomycin therapy. Int J Med Microbiol. 2015;305(8):902–9. doi: 10.1016/j.ijmm.2015.09.005. [DOI] [PubMed] [Google Scholar]
  • 95.Paulsen IT, Banerjei L, Myers GSA, Nelson KE, R Seshadri R, Read TD, Fouts DE, Eisen JA, Gill SR, Heidelberg JF. Role of mobile DNA in the evolution of vancomycin-resistant Enterococcus faecalis. Science. 2003;299(5615):2071–2074. doi: 10.1126/science.1080613. [DOI] [PubMed] [Google Scholar]
  • 96.Zheng B, Tomita H, Inoue T, Ike Y. Isolation of vanB-type Enterococcus faecalis strains from nosocomial infections: First report of the isolation and identification of the pheromone-responsive plasmids pmg2200, encoding vanB-type vancomycin resistance and a bac41-type bacteriocin, and pmg2201, encoding erythromycin resistance and cytolysin (hly/bac) Antimicrob Agents Chemother. 2009;53(2):735–47. doi: 10.1128/AAC.00754-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Heaton MP, Discotto LF, Pucci MJ, Handwerger S. Mobilization of vancomycin resistance by transposon-mediated fusion of a vanA plasmid with an Enterococcus faecium sex pheromone-response plasmid. Gene. 1996;171(1):9–17. doi: 10.1016/0378-1119(96)00022-4. [DOI] [PubMed] [Google Scholar]
  • 98.Manson JM, Hancock LE, Gilmore MS. Mechanism of chromosomal transfer of Enterococcus faecalis pathogenicity island, capsule, antimicrobial resistance, and other traits. Proc Natl Acad Sci USA. 2010;107(27):12269–74. doi: 10.1073/pnas.1000139107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Arthur M, Molinas C, Depardieu F, Courvalin P. Characterization of Tn1546, a Tn3-related transposon conferring glycopeptide resistance by synthesis of depsipeptide peptidoglycan precursors in Enterococcus faecium BM4147. J Bacteriol. 1993;175(1):117–27. doi: 10.1128/jb.175.1.117-127.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Carias LL, Rudin SD, Donskey CJ, Rice LB. Genetic linkage and cotransfer of a novel, vanB-containing transposon (tn5382) and a low-affinity penicillin-binding protein 5 gene in a clinical vancomycin-resistant Enterococcus faecium isolate. J Bacteriol. 1998;180(17):4426–4434. doi: 10.1128/jb.180.17.4426-4434.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Dahl KH, Lundblad EW, Rokenes TP, Olsvik O, Sundsfjord A. Genetic linkage of the vanB2 gene cluster to tn5382 in vancomycin-resistant Enterococci and characterization of two novel insertion sequences. Microbiology. 2000;146(Pt 6):1469–79. doi: 10.1099/00221287-146-6-1469. [DOI] [PubMed] [Google Scholar]
  • 102.Garnier F, Taourit S, Glaser P, Courvalin P, Galimand M. Characterization of transposon tn1549, conferring vanB-type resistance in Enterococcus spp. Microbiology. 2000;146(6):1481–9. doi: 10.1099/00221287-146-6-1481. [DOI] [PubMed] [Google Scholar]
  • 103.Rice LB, Lenore L, Carias LL, Marshall S, Rudin SD, Hutton-Thomas R. Tn5386, a novel tn916-like mobile element in Enterococcus faecium D344R that interacts with tn916 to yield a large genomic deletion. J Bacteriol. 2005;187(19):6668–77. doi: 10.1128/JB.187.19.6668-6677.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Palmer KL, van Schaik W, Willems RJL, et al. Enterococcal Genomics. In: Gilmore MS, Clewell DB, Ike Y, et al., editors. Enterococci: From Commensals to Leading Causes of Drug Resistant Infection [Internet] Boston: Massachusetts Eye and Ear Infirmary; 2014. 2014-. Available from: https://www.ncbi.nlm.nih.gov/books/NBK190425/ [PubMed] [Google Scholar]
  • 105.van Schaik W, Willems RJL. Genome-based insights into the evolution of Enterococci. Clin Microbiol Infect. 2010;16(6):527–32. doi: 10.1111/j.1469-0691.2010.03201.x. [DOI] [PubMed] [Google Scholar]
  • 106.Qin X, et al. Complete genome sequence of Enterococcus faecium strain TX16 and comparative genomic analysis of Enterococcus faecium genomes. BMC Microbiol. 2012;12:135. doi: 10.1186/1471-2180-12-135. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Palmer KL, et al. Comparative genomics of Enterococci: Variation in Enterococcus faecalis, clade structure in E. faecium, and defining characteristics of E. gallinarum and E. casseliflavus. MBio. 2012;3(1):e00318–11. doi: 10.1128/mBio.00318-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Palmer KL, Gilmore MS. Multidrug-resistant Enterococci lack CRISPR-cas. MBio. 2010;1(4):e00227–10. doi: 10.1128/mBio.00227-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.van Schaik W, et al. Pyrosequencing-based comparative genome analysis of the nosocomial pathogen Enterococcus faecium and identification of a large transferable pathogenicity island. BMC Genomics. 2010;11:239. doi: 10.1186/1471-2164-11-239. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Leavis HL, Willems RJL, van Wamel WJB, Schuren FH, Caspers MPM, Bonten MJM. Insertion sequence-driven diversification creates a globally dispersed emerging multiresistant subspecies of E. faecium. PLoS Pathog. 2007;3(1):e7. doi: 10.1371/journal.ppat.0030007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Willems RJ, van Schaik W. Transition of Enterococcus faecium from commensal organism to nosocomial pathogen. Future Microbiol. 2009;4(9):1125–1135. doi: 10.2217/fmb.09.82. [DOI] [PubMed] [Google Scholar]
  • 112.Werner G, Fleige C, Geringer U, van Schaik W, Klare I, Witte W. IS element IS16 as a molecular screening tool to identify hospital-associated strains of Enterococcus faecium. BMC Infect Dis. 2011;11:80. doi: 10.1186/1471-2334-11-80. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Ruiz-Garbajosa P, et al. Multilocus sequence typing scheme for Enterococcus faecalis reveals hospital-adapted genetic complexes in a background of high rates of recombination. J Clin Microbiol. 2006;44(6):2220–8. doi: 10.1128/JCM.02596-05. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Bourgogne A, et al. Large scale variation in Enterococcus faecalis illustrated by the genome analysis of strain OG1RF. Genome Biol. 2008;9(7):R110. doi: 10.1186/gb-2008-9-7-r110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Galloway-Peña JR, Roh JH, Latorre M, Qin X, Murray BE. Genomic and SNP analyses demonstrate a distant separation of the hospital and community-associated clades of Enterococcus faecium. PLoS One. 2012;7(1):e30187. doi: 10.1371/journal.pone.0030187. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Willems RJL, Top J, van Schaik W, Leavis H, Bonten M, Sirén J, Hanage WP, Corander J. Restricted gene flow among hospital subpopulations of Enterococcus faecium. MBio. 2012;3(4):e00151–12. doi: 10.1128/mBio.00151-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Lebreton F, et al. Emergence of epidemic multidrug-resistant Enterococcus faecium from animal and commensal strains. MBio. 2013;4(4):e00534–13. doi: 10.1128/mBio.00534-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.de Regt MJA, van Schaik W, van Luit-Asbroek M, Dekker HAT, van Duijkeren E, Koning CJM, Bonten MJM, Willems RJL. Hospital and community ampicillin-resistant Enterococcus faecium are evolutionarily closely linked but have diversified through niche adaptation. PLoS One. 2012;7(2):e30319. doi: 10.1371/journal.pone.0030319. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Woodford N, Adebiyi AM, Palepou MF, Cookson BD. Diversity of vanA glycopeptide resistance elements in Enterococci from humans and nonhuman sources. Antimicrob Agents Chemother. 1998;42(3):502–8. doi: 10.1128/aac.42.3.502. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Willems RJ, Top J, van den Braak N, van Belkum A, Mevius DJ, Hendriks G, van Santen-Verheuvel M, van Embden JD. Molecular diversity and evolutionary relationships of tn1546-like elements in Enterococci from humans and animals. Antimicrob Agents Chemother. 1999;43(3):483–91. doi: 10.1128/aac.43.3.483. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Lester CH, Frimodt-Møller N, Sørensen TL, Monnet DL, Hammerum AM. In vivo transfer of the vanA resistance gene from an Enterococcus faecium isolate of animal origin to an E. faecium isolate of human origin in the intestines of human volunteers. Antimicrob Agents Chemother. 2006;50(2):596–9. doi: 10.1128/AAC.50.2.596-599.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Kuhn I, et al. Occurrence and relatedness of vancomycin-resistant Enterococci in animals, humans, and the environment in different European regions. Appl Environ Microbiol. 2005;71(9):5383–90. doi: 10.1128/AEM.71.9.5383-5390.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Maiden MC, et al. Multilocus sequence typing: A portable approach to the identification of clones within populations of pathogenic microorganisms. Proc Natl Acad Sci USA. 1998;95(6):3140–5. doi: 10.1073/pnas.95.6.3140. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Homan WL, Tribe D, Poznanski S, Li M, Hogg G, Spalburg E, van Embden JDA, Willems RJL. Multilocus sequence typing scheme for Enterococcus faecium. J Clin Microbiol. 2002;40(6):1963–1971. doi: 10.1128/JCM.40.6.1963-1971.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Nallapareddy SR, Wenxiang H, Weinstock GM, Murray BE. Molecular characterization of a widespread, pathogenic, and antibiotic resistance-receptive Enterococcus faecalis lineage and dissemination of its putative pathogenicity island. J Bacteriol. 2005;187(16):5709–18. doi: 10.1128/JB.187.16.5709-5718.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Leavis HL, Bonten MJM, Willems RJL. Identification of high-risk Enterococcal clonal complexes: Global dispersion and antibiotic resistance. Curr Opin Microbiol. 2006;9(5):454–60. doi: 10.1016/j.mib.2006.07.001. [DOI] [PubMed] [Google Scholar]
  • 127.McBride SM, Fischetti VA, Leblanc DJ, Moellering RC, Gilmore MS. Genetic diversity among Enterococcus faecalis. PLoS One. 2007;2(7):e582. doi: 10.1371/journal.pone.0000582. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Turner KME, Hanage WP, Fraser C, Connor TR, Spratt BG. Assessing the reliability of eBURST using simulated populations with known ancestry. BMC Microbiol. 2007;7:30. doi: 10.1186/1471-2180-7-30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Devriese LA, Van de Kerckhove A, Kilpper-Bälz R, Schleifer KH. Characterization and identification of Enterococcus species isolated from the intestines of animals. International Journal of Systematic and Evolutionary Microbiology. 1987;37(3):257–259. [Google Scholar]
  • 130.Willems RJL, Top J, van Belkum A, Endtz H, Mevius D, Stobberingh E, van den Bogaard A, van Embden JDA. Host specificity of vancomycin-resistant Enterococcus faecium. Journal of Infectious Diseases. 2000;182(3):816–823. doi: 10.1086/315752. [DOI] [PubMed] [Google Scholar]
  • 131.Ostrolenk M, Kramer N, Cleverdon RC. Comparative studies of Enterococci and Escherichia coli as indices of pollution. J Bacteriol. 1947;53(2):197–203. doi: 10.1128/jb.53.2.197-203.1947. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Harwood VJ, Whitlock J, Withington V. Classification of antibiotic resistance patterns of indicator bacteria by discriminant analysis: Use in predicting the source of fecal contamination in subtropical waters. Appl Environ Microbiol. 2000;66(9):3698–704. doi: 10.1128/aem.66.9.3698-3704.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Boehm AB, Sassoubre LM. Enterococci as Indicators of Environmental Fecal Contamination. In: Gilmore MS, Clewell DB, Ike Y, et al., editors. Enterococci: From Commensals to Leading Causes of Drug Resistant Infection [Internet] Boston: Massachusetts Eye and Ear Infirmary; 2014. 2014-. Available from: https://www.ncbi.nlm.nih.gov/books/NBK190421/ [PubMed] [Google Scholar]
  • 134.Sinclair JL, Alexander M. Role of resistance to starvation in bacterial survival in sewage and lake water. Appl Environ Microbiol. 1984;48(2):410–5. doi: 10.1128/aem.48.2.410-415.1984. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Leclercq R, Kenny Oberlé K, Galopin S, Cattoir V, Budzinski H, Petit F. Changes in Enterococcal populations and related antibiotic resistance along a medical center-wastewater treatment plant-river continuum. Appl Environ Microbiol. 2013;79(7):2428–34. doi: 10.1128/AEM.03586-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.van Embden JD, Engel HW, van Klingeren B. Drug resistance in group D Streptococci of clinical and nonclinical origin: Prevalence, transferability, and plasmid properties. Antimicrob Agents Chemother. 1977;11(6):925–32. doi: 10.1128/aac.11.6.925. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Wade TJ, Nitika Pai N, Eisenberg JNS, Colford JM. Do U.S. Environmental protection agency water quality guidelines for recreational waters prevent gastrointestinal illness? A systematic review and meta-analysis. Environ Health Perspect. 2003;111(8):1102–9. doi: 10.1289/ehp.6241. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Chenoweth C, Schaberg D. The epidemiology of Enterococci. European Journal of Clinical Microbiology and Infectious Diseases. 1990;9:80–89. doi: 10.1007/BF01963631. [DOI] [PubMed] [Google Scholar]
  • 139.Noble CJ. Carriage of group D Streptococci in the human bowel. Journal of Clinical Pathology. 1978;31(12):1182–1186. doi: 10.1136/jcp.31.12.1182. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Layton BA, Walters SP, Lam LH, Boehm AB. Enterococcus species distribution among human and animal hosts using multiplex PCR. J Appl Microbiol. 2010;109(2):539–47. doi: 10.1111/j.1365-2672.2010.04675.x. [DOI] [PubMed] [Google Scholar]
  • 141.Hooper LV, Midtvedt T, Gordon JI. How host-microbial interactions shape the nutrient environment of the mammalian intestine. Annu Rev Nutr. 2002;22:283–307. doi: 10.1146/annurev.nutr.22.011602.092259. [DOI] [PubMed] [Google Scholar]
  • 142.Littman DR, Pamer EG. Role of the commensal microbiota in normal and pathogenic host immune responses. Cell Host Microbe. 2011;10(4):311–23. doi: 10.1016/j.chom.2011.10.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Are A, Aronsson L, Wang S, Greicius G, Lee YK, Gustafsson J, Pettersson S, Arulampalam V. Enterococcus faecalis from newborn babies regulate endogenous PPAR-gamma activity and IL-10 levels in colonic epithelial cells. Proc Natl Acad Sci USA. 2008;105(6):1943–8. doi: 10.1073/pnas.0711734105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Wang S. Infant intestinal Enterococcus faecalis down-regulates inflammatory responses in human intestinal cell lines. World Journal of Gastroenterology. 2008;14(7):1067. doi: 10.3748/wjg.14.1067. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Chen C, et al. Probiotics have clinical, microbiologic, and immunologic efficacy in acute infectious diarrhea. The Pediatric Infectious Disease Journal. 2010;29(2):135–138. doi: 10.1097/inf.0b013e3181b530bf. [DOI] [PubMed] [Google Scholar]
  • 146.Turnbaugh PJ, Ley RE, Hamady M, Fraser-Liggett CM, Knight R, Gordon JI. The human microbiome project. Nature. 2007;449(7164):804–10. doi: 10.1038/nature06244. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Human Microbiome Project Consortium. Structure, function and diversity of the healthy human microbiome. Nature. 2012;486(7402):207–14. doi: 10.1038/nature11234. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Arumugam M, et al. Enterotypes of the human gut microbiome. Nature. 2011;473(7346):174–80. doi: 10.1038/nature09944. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Koren O, Knights D, Gonzalez A, Waldron L, Segata N, Knight R, Huttenhower C, Ley RE. A guide to enterotypes across the human body: Meta-analysis of microbial community structures in human microbiome datasets. PLoS Comput Biol. 2013;9(1):e1002863. doi: 10.1371/journal.pcbi.1002863. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Ding T, Schloss PD. Dynamics and associations of microbial community types across the human body. Nature. 2014;509(7500):357–60. doi: 10.1038/nature13178. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Donskey CJ, Hanrahan JA, Hutton RA, Rice LB. Effect of parenteral antibiotic administration on persistence of vancomycin-resistant Enterococcus faecium in the mouse gastrointestinal tract. J Infect Dis. 1999;180(2):384–90. doi: 10.1086/314874. [DOI] [PubMed] [Google Scholar]
  • 152.Sjolund M, Wreiber K, Andersson DI, Blaser MJ, Engstrand L. Long-term persistence of resistant Enterococcus species after antibiotics to eradicate Helicobacter pylori. Ann Intern Med. 2003;139(6):483–7. doi: 10.7326/0003-4819-139-6-200309160-00011. [DOI] [PubMed] [Google Scholar]
  • 153.van der Waaij D, Berghuis-de Vries JM, Lekkerkerk Lekkerkerk-v. Colonization resistance of the digestive tract in conventional and antibiotic-treated mice. J Hyg (Lond) 1971;69(3):405–11. doi: 10.1017/s0022172400021653. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Freter R. The fatal enteric cholera infection in the guinea pig, achieved by inhibition of normal enteric flora. J Infect Dis. 1955;97(1):57–65. doi: 10.1093/infdis/97.1.57. [DOI] [PubMed] [Google Scholar]
  • 155.Bohnhoff C, Miller CP, Martin WP. Resistance of the mouse's intestinal tract to experimental Salmonella infection. J Exp Med. 1964;120:817–28. doi: 10.1084/jem.120.5.817. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Ubeda C, et al. Intestinal microbiota containing Barnesiella species cures vancomycin-resistant Enterococcus faecium colonization. Infect Immun. 2013;81(3):965–73. doi: 10.1128/IAI.01197-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Caballero S, Kim S, Carter RA, Leiner IM, Sušac B, Miller L, Kim GJ, Lilan L, Pamer EG. Cooperating commensals restore colonization resistance to vancomycin-resistant Enterococcus faecium. Cell Host Microbe. 2017;21(5):592–602. doi: 10.1016/j.chom.2017.04.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Cash HL, Whitham CV, Behrendt CL, Hooper LV. Symbiotic bacteria direct expression of an intestinal bactericidal lectin. Science. 2006;313(5790):1126–30. doi: 10.1126/science.1127119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Brandl K, Plitas G, Schnabl B, DeMatteo RP, Pamer EG. MyD88-mediated signals induce the bactericidal lectin RegIIIγ and protect mice against intestinal Listeria monocytogenes infection. J Exp Med. 2007;204(8):1891–1900. doi: 10.1084/jem.20070563. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Vaishnava S, Behrendt CL, Ismail AS, Eckmann L, Hooper LV. Paneth cells directly sense gut commensals and maintain homeostasis at the intestinal host-microbial interface. Proc Natl Acad Sci USA. 2008;105(52):20858–63. doi: 10.1073/pnas.0808723105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Mukherjee S, et al. Antibacterial membrane attack by a pore-forming intestinal c-type lectin. Nature. 2013;505(7481):103–7. doi: 10.1038/nature12729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Brandl K, Plitas G, Mihu CN, Ubeda C, Jia T, Fleisher M, Schnabl B, DeMatteo RP, Pamer EG. Vancomycin-resistant Enterococci exploit antibiotic-induced innate immune deficits. Nature. 2008;455(7214):804–807. doi: 10.1038/nature07250. Web. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Kinnebrew MA, Ubeda C, Zenewicz LA, Smith N, Flavell RA, Pamer EG. Bacterial flagellin stimulates toll-like receptor 5-dependent defense against vancomycin-resistant Enterococcus infection. J Infect Dis. 2010;201(4):534–543. doi: 10.1086/650203. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Godl K, Johansson MEV, Lidell ME, Mörgelin M, Karlsson H, Olson FJ, Gum JR, Kim YS, Hansson GC. The N terminus of the MUC2 mucin forms trimers that are held together within a trypsin-resistant core fragment. J Biol Chem. 2002;277(49):47248–56. doi: 10.1074/jbc.M208483200. [DOI] [PubMed] [Google Scholar]
  • 165.Donskey CJ. The role of the intestinal tract as a reservoir and source for transmission of nosocomial pathogens. Clin Infect Dis. 2004;39(2):219–26. doi: 10.1086/422002. [DOI] [PubMed] [Google Scholar]
  • 166.Johansson ME, Jakobsson HE, Holmén-Larsson J, Schütte A, Ermund A, Rodríguez-Piñeiro AM, Arike L, Wising C, Svensson F, Bäckhed F, Hansson GC. Normalization of host intestinal mucus layers requires long-term microbial colonization. Cell Host Microbe. 2015;18(5):582–92. doi: 10.1016/j.chom.2015.10.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Petersson J, Schreiber O, Hansson GC, Gendler SJ, Velcich A, Lundberg JO, Roos S, Holm L, Phillipson M. Importance and regulation of the colonic mucus barrier in a mouse model of colitis. Am J Physiol Gastrointest Liver Physiol. 2011;300(2):G327–33. doi: 10.1152/ajpgi.00422.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Frantz AL, Rogier EW, Weber CR, Shen L, Cohen DA, Fenton LA, Bruno MEC, Kaetzel CS. Targeted deletion of MyD88 in intestinal epithelial cells results in compromised antibacterial immunity associated with downregulation of polymeric immunoglobulin receptor, mucin-2, and antibacterial peptides. Mucosal Immunol. 2012;5(5):501–12. doi: 10.1038/mi.2012.23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Wlodarska M, Willing B, Keeney KM, Menendez A, Bergstrom KS, Gill N, Russell SL, Vallance BA, Finlay BB. Antibiotic treatment alters the colonic mucus layer and predisposes the host to exacerbated Citrobacter rodentium-induced colitis. Infect Immun. 2011;79(4):1536–45. doi: 10.1128/IAI.01104-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Johansson MEV, Phillipson M, Petersson J, Velcich A, Holm L, Hansson GC. The inner of the two muc2 mucin-dependent mucus layers in colon is devoid of bacteria. Proc Natl Acad Sci USA. 2008;105(39):15064–9. doi: 10.1073/pnas.0803124105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Vaishnava S, Miwako Yamamoto M, Severson KM, Ruhn KA, Yu X, Koren O, Ley R, Wakeland EK, Hooper LV. The antibacterial lectin RegIIIγ promotes the spatial segregation of microbiota and host in the intestine. Science. 2011;334(6053):255–8. doi: 10.1126/science.1209791. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Loonen LMP, Stolte EH, Jaklofsky MTJ, Meijerink M, Dekker J, van Baarlen P, Wells JM. RegIIIγ-deficient mice have altered mucus distribution and increased mucosal inflammatory responses to the microbiota and enteric pathogens in the ileum. Mucosal Immunol. 2014;7(4):939–47. doi: 10.1038/mi.2013.109. [DOI] [PubMed] [Google Scholar]
  • 173.Caballero S, Carter R, Ke X, Boze S, Leiner IM, Kim GJ, Miller L, Ling L, Manova K, Pamer EG. Distinct but spatially overlapping intestinal niches for vancomycin-resistant Enterococcus faecium and carbapenem-resistant Klebsiella pneumoniae. PLoS Pathog. 2015;11(9):e1005132. doi: 10.1371/journal.ppat.1005132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Borrero J, Chen Y, Dunny GM, Kaznessis YN. Modified lactic acid bacteria detect and inhibit multiresistant enterococci. ACS Synthetic Biology. 2015;4(3):299–30. doi: 10.1021/sb500090b. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Millette M, Cornut G, Dupont C, Shareck F, Archambault D, Lacroix M. Capacity of human nisin- and pediocin-producing lactic acid bacteria to reduce intestinal colonization by vancomycin-resistant enterococci. Appl Environ Microbiol. 2008;74(7):1997–2003. doi: 10.1128/AEM.02150-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Weisser M, Oostdijk EA, Willems RJL, Bonten MJM, Frei R, Elzi L, Halter J, Widmer AF, Top J. Dynamics of ampicillin-resistant Enterococcus faecium clones colonizing hospitalized patients: Data from a prospective observational study. BMC Infect Dis. 2012;12:68. doi: 10.1186/1471-2334-12-68. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Montealegre MC, Singh KV, Murray BE. Gastrointestinal tract colonization dynamics by different Enterococcus faecium clades. J Infect Dis. 2016;213(12):1914–22. doi: 10.1093/infdis/jiv597. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178.Kommineni Sushma, Bretl Daniel J, Lam Vy, Chakraborty Rajrupa, Hayward Michael, Simpson Pippa, Cao Yumei, Bousounis Pavlos, Kristich Christopher J, Salzman Nita H. Bacteriocin production augments niche competition by enterococci in the mammalian gastrointestinal tract. Nature. 2015;526(7575):719–22. doi: 10.1038/nature15524. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Gilmore MS, Rauch M, Ramsey MM, Himes PR, Varahan S, Manson JM, Lebreton F, Hancock LE. Pheromone killing of multidrug-resistant Enterococcus faecalis V583 by native commensal strains. Proc Natl Acad Sci USA. 2015;112(23):7273–8. doi: 10.1073/pnas.1500553112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180.Duerkop BA, Clements CV, Rollins D, Rodrigues JLM, Hooper LV. A composite bacteriophage alters colonization by an intestinal commensal bacterium. Proc Natl Acad Sci USA. 2012;109(43):17621–6. doi: 10.1073/pnas.1206136109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.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):204–10. doi: 10.1128/IAI.70.1.204-210.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Franz C, Huch M, Abriouel H, Holzapfel W, Gálvez A. Enterococci as probiotics and their implications in food safety. Int J Food Microbiol. 2011;151(2):125–40. doi: 10.1016/j.ijfoodmicro.2011.08.014. [DOI] [PubMed] [Google Scholar]
  • 183.Domann E, Hain T, Ghai R, Billion A, Kuenne C, Zimmermann K, Chakraborty T. Comparative genomic analysis for the presence of potential Enterococcal virulence factors in the probiotic Enterococcus faecalis strain symbioflor 1. Int J Med Microbiol. 2007;297(7–8):533–9. doi: 10.1016/j.ijmm.2007.02.008. [DOI] [PubMed] [Google Scholar]
  • 184.Allen SJ, Martinez EG, Gregorio GV, Dans LF. Probiotics for treating acute infectious diarrhoea. Cochrane Database Syst Rev. 2010;(11):CD003048. doi: 10.1002/14651858.CD003048.pub3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Bellomo G, Mangiagle A, Nicastro L, Frigerio G. A controlled double-blind study of SF68 strain as a new biological preparation for the treatment of diarrhoea in pediatrics. Current Therapeutic Research. 1980;28:927–934. [Google Scholar]
  • 186.Buydens P, Debeuckelaere S. Efficacy of SF68 in the treatment of acute diarrhoea. A placebo-controlled trial. Scandinavian Journal of Gastroenterology. 1996;31:887–891. doi: 10.3109/00365529609051997. [DOI] [PubMed] [Google Scholar]
  • 187.van Nood E, et al. Duodenal infusion of donor feces for recurrent Clostridium difficile. N Engl J Med. 2013;368(5):407–15. doi: 10.1056/NEJMoa1205037. [DOI] [PubMed] [Google Scholar]
  • 188.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. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 189.Lahti L, Salojärvi J, Salonen A, Scheffer M, de Vos WM. Tipping elements in the human intestinal ecosystem. Nat Commun. 2014;5:4344. doi: 10.1038/ncomms5344. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Buffie CG, et al. Precision microbiome reconstitution restores bile acid mediated resistance to Clostridium difficile. Nature. 2014 doi: 10.1038/nature13828. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Manley KJ, Fraenkel MB, Mayall BC, Power DA. Probiotic treatment of vancomycin-resistant enterococci: A randomised controlled trial. Med J Aust. 2007;186(9):454–7. doi: 10.5694/j.1326-5377.2007.tb00995.x. [DOI] [PubMed] [Google Scholar]
  • 192.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):872–7. doi: 10.1097/MCG.0b013e318227439f. [DOI] [PubMed] [Google Scholar]
  • 193.Vidal M, Forestier C, Charbonnel N, Henard S, Rabaud C, Lesens O. Probiotics and intestinal colonization by vancomycin-resistant enterococci in mice and humans. J Clin Microbiol. 2010;48(7):2595–8. doi: 10.1128/JCM.00473-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Doron S, Hibberd PL, Goldin B, Thorpe C, McDermott L, Snydman DR. Effect of lactobacillus rhamnosus GG administration on vancomycin-resistant enterococcus colonization in adults with comorbidities. Antimicrob Agents Chemother. 2015;59(8):4593–9. doi: 10.1128/AAC.00300-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.de Regt MJA, Willems RJL, Hené RJ, Siersema PD, Verhaar HJJ, Hopmans TEM, Bonten MJM. Effects of probiotics on acquisition and spread of multiresistant Enterococci. Antimicrob Agents Chemother. 2010;54(7):2801–5. doi: 10.1128/AAC.01765-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.Stein RR, von Mering C, Bucci V, Toussaint NC, Buffie GC, Rätsch G, Pamer EG, Sander C, Xavier JB. Ecological modeling from time-series inference: Insight into dynamics and stability of intestinal microbiota. PLoS Comput Biol. 2013;9(12):e1003388. doi: 10.1371/journal.pcbi.1003388. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Khanna S, Pardi DS, Kelly CR, Kraft CS, Dhere T, Henn MR, Lombardo MJ, Vulic M, Ohsumi T, Winkler J, Pindar C, McGovern BH, Pomerantz RJ, Aunins JG, Cook DN, Hohmann EL. A novel microbiome therapeutic increases gut microbial diversity and prevents recurrent clostridium difficile infection. J Infect Dis. 2016;214(2):173–81. doi: 10.1093/infdis/jiv766. [DOI] [PubMed] [Google Scholar]
  • 198.Ratner M. Seres’s pioneering microbiome drug fails mid-stage trial. Nat Biotech. 2016;34:1004–5. doi: 10.1038/nbt1016-1004b. [DOI] [PubMed] [Google Scholar]
  • 199.Pamer EG. Fecal microbiota transplantation: Effectiveness, complexities, and lingering concerns. Mucosal Immunol. 2014;7(2):210–4. doi: 10.1038/mi.2013.117. [DOI] [PubMed] [Google Scholar]
  • 200.Stiefel U, Nerandzic MM, Pultz MJ, Donskey CJ. Gastrointestinal colonization with a cephalosporinase-producing Bacteroides species preserves colonization resistance against vancomycin-resistant Enterococcus and Clostridium difficile in cephalosporin-treated mice. Antimicrob Agents Chemother. 2014;58(8):4535–4542. doi: 10.1128/AAC.02782-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Zeevi D, Korem T, Zmora N, Israeli D, Rothschild D, Weinberger A, Ben-Yacov O, Lador D, Avnit-Sagi T, Lotan-Pompan M, Suez J, Mahdi JA, Matot E, Malka G, Kosower N, Rein M, Zilberman-Schapira G, Dohnalová L, Pevsner-Fischer M, Bikovsky R, Halpern Z, Elinav E, Segal E. Personalized nutrition by prediction of glycemic responses. Cell. 2015;163(5):1079–94. doi: 10.1016/j.cell.2015.11.001. [DOI] [PubMed] [Google Scholar]

RESOURCES