Multiple hypervirulent methicillin-sensitive Staphylococcus aureus lineages contribute towards poor patient outcomes in orthopedic device-related infections ============================================================================================================================================================ * Virginia Post * Ben Pascoe * Evangelos Mourkas * Jessica K. Calland * Matthew D. Hitchings * Christoph Erichsen * Julian Fischer * Mario Morgenstern * R. Geoff Richards * Samuel K. Sheppard * T. Fintan Moriarty ## Abstract Staphylococci are the most common cause of orthopedic device-related infections (ODRIs), with *Staphylococcus aureus* responsible for a third or more of cases. This prospective clinical and laboratory study investigated the association of genomic and phenotypic variation with treatment outcomes in ODRI isolates. Eighty-six invasive *S. aureus* isolates were collected from patients with ODRI, and clinical outcome was assessed after a follow-up examination of 24 months. Each patient was then considered to have been “cured” or “not cured” based on predefined clinical criteria. Whole genome sequencing and molecular characterization identified isolates belonging to globally circulating community- and hospital-acquired pandemic lineages. Most isolates were phenotypically susceptible to methicillin and lacked the SCC*mec* cassette (MSSA), but contained several (hyper) virulence genes, including toxins and biofilm genes. While recognizing the role of the host immune response, we identify characteristics of isolate genomes that, with larger datasets, could help contribute to infection severity or clinical outcome predictions. While this and several other studies reinforce the role antibiotic resistance (e.g., MRSA infection) has on treatment failure, it is important not to overlook MSSA that can cause equally destructive infections and lead to poor patient outcomes. **Importance** *Staphylococcus aureus* is a prominent cause of orthopedic device-associated infections, yet little is known about how the infecting pathogen, and specifically the repertoire of genome-encoded virulence factors can impact treatment outcome. Past studies have focused on distinguishing commensal from invasive *S. aureus* isolates but in this study, we aim to investigate traits in infecting isolates that influence patient outcomes. Invasive *S. aureus* isolates were collected from orthopedic-device related infection patients and categorized according to the success of subsequent treatment (“cured” /”not cured”), as determined following hospital discharge two years after initial presentation. Several MSSA hypervirulent clones were associated with a “not cured” clinical outcome. Improved understanding of the bacterial traits associated with treatment failure in ODRI will inform the risk assessment, prognosis, and therapy of these infections. Keywords * *Staphylococcus aureus* * MRSA * MSSA * virulence factors * antibiotic resistance * orthopedic device-related infections ## Introduction The most challenging complication in orthopedic surgery is orthopedic-device related infection (ODRI), with incidence ranging from 0.7 % to 4.2 % for elective orthopedic surgeries (1–4). Incidence increases to over 30 % following operative fixation of complex open fractures (5, 6). While patient health is a key risk factor (including a high BMI and chronic immunosuppression) for poor treatment outcomes (7, 8), there is evidence that pathogen genetic diversity can be an indicator of patient outcome (9–12). *Staphylococcus aureus* is the most common infecting agent (2, 3, 13–15) and treatment outcome is often complicated by infection with antimicrobial resistant lineages e.g., methicillin-resistant *S. aureus* (MRSA) (8, 16). Combined with high virulence potential, MRSA are difficult to treat and are a global healthcare concern. Despite this, predicting treatment outcomes based on bacterial phenotypes/genotypes remains difficult (8, 10, 11, 17). Hypervirulent pandemic clonal lineages have helped spread *S. aureus* around the world and the geographic distribution of many lineages is dynamic, with different lineages dominating infections in specific global regions (18, 19). Waves of MRSA lineages have risen and been replaced since the emergence of MRSA in the 1940s (18, 20). These highly structured populations can be grouped into clonal complexes (CCs) that share five or more alleles at seven multi-locus sequence typing (MLST) loci (21–23). Community-associated MRSA (CA-MRSA) have begun to replace hospital-associated MRSA (HA-MRSA) as the dominant epidemic strains (24). The most prevalent lineages include 5 global community acquired (CA-MRSA) genotypes CC1, CC8, CC30, CC59 and CC80. The most common hospital-acquired lineages are CC5, CC22 (UK), CC239 and CC45 (21–23). Combining MLST with traditional molecular typing techniques such as identification of the staphylococcal cassette chromosome *mec* (SCC*mec*) (25) and *spa* repeat regions (26) can provide a nomenclature to describe relevant epidemic clones can be provided. Lineages can acquire advantageous traits, such as antibiotic resistance, which proliferate in the population through the decedents of successful strains. This likely occurs in many instances, however, the extent of this in the context of ODRI remains to be determined. Despite the growing concern of MRSA lineages, MSSA isolates are often the most common in invasive surgery-related infections (27, 28). Convergent evolution in several CA-lineages potentially balances the fitness costs of expressing AMR genes with the acquisition of multiple virulence factors (29–32). Specific genes encoding putative virulence factors in invasive *S. aureus* disease involve evasion of immune defenses, including the ability to adhere to and invade host tissues - essential for ODRI (33, 34). A large body of work has identified many virulence factors, including Microbial Surface Components Recognizing Adhesive Matrix Molecules (MSCRAMMs), the polysaccharide intercellular adhesion (PIA), the Staphylococcal protein A, extracellular proteins such as coagulase, Staphylococcal enterotoxins (SE), exfoliatins (ET), toxic shock syndrome toxin (TSST), staphyloxanthin, hemolysins and Panton Valentine leukocidin (PVL) that have a crucial impact on the pathogenicity of *S. aureus* infections (35–38). Genome-wide association studies have been applied to numerous bacterial species (39, 40), and identified genes or genetic elements associated with disease that transcend clonal patterns of inheritance (not confined to specific lineages) in Staphylococci. Phenotype filtering techniques have attempted to assess patient risk and predict disease outcome from genetic data based on enrichment for disease-associated traits (12, 41–43). GWAS approaches have shown promising disease prediction results (31, 43, 44), as well as AMR profiles (45). As in other bacterial species (46–48), a better understanding of genome and transcriptome variation in hypervirulent infection types shows promise for our ability to predict disease severity in Staphylococci (49). In this study we investigate the association between treatment outcomes determined 2 years post-operatively in patients with *S. aureus* ODRI and phenotypic and genotypic features of the infecting pathogen. Building on recent studies aiming to predict *S. aureus* virulence from genome sequence (43, 44), we aim to distinguish high risk lineages, isolates and genes. These features were correlated with mortality rate in a simplified virulence model using *Galleria mellonella*. ## Results *S. aureus* isolates were collected from 86 patients undergoing operative revision of an ODRI (**Supplementary table S1**). Patient outcomes were assessed after a 2-year follow-up period and “cured” patients were free of infection, surgical and systemic antibiotic therapy had ceased with function of the affected joint or limb restored. If one or more of these parameters was negative, patients were considered to have had a “not cured” outcome (8, 12). Most patients enrolled in the study received successful treatment and were among the “cured” cohort (**Table 1**; n= 65/86; 75.6%). Treatment was unsuccessful in 21 patients (n=21/86; 24.4% “not-cured”) and multiple revision surgeries were necessary for nearly all patients (n=83/86; 96.5%). View this table: [Table 1:](http://medrxiv.org/content/early/2022/10/22/2022.10.21.22280349/T1) Table 1: Cohort description ### Host-associated risk factors Extensive patient data, types of implanted devices and clinical presentation were recorded for each infected patient (**Table 1**). Additionally, the effects of patient co-morbidities (such as diabetes or obesity), fracture types or the time of symptoms onset on treatment outcome were analyzed (**Table 2**). None of these prognostic factors alone significantly decreased cure rate. View this table: [Table 2:](http://medrxiv.org/content/early/2022/10/22/2022.10.21.22280349/T2) Table 2: Patient risk factors ### Multiple lineages contribute to poor patient outcome A maximum-likelihood phylogeny was constructed based on shared coding sequences, present in 95% or more isolates (**Figure 1A; Supplementary table S1**). The collected isolates were genetically diverse and represented 19 different sequence types (STs) (**Figure 1B**) based on the 7 loci scheme for *S. aureus* and could be grouped into 18 clonal complexes (CCs) based on 5 or more shared MLST loci (21). No clear clustering was observed between “cured” and “not-cured” isolates. Consistent with other European surveillance efforts, six of our seven most common lineages matched common pandemic *S. aureus* lineages (CC5, CC8, CC22, CC30, CC45 and CC59) previously identified by the ESCMID study group (collected from 26 European countries between 2006-7) (50). Our isolates predominantly lacked the *SCCmec* cassette, which confers methicillin resistance and were classified as methicillin-susceptible *S. aureus* (MSSA; n= 81/86; 94%) (**Figure 1C**). ![Figure 1:](http://medrxiv.org/https://www.medrxiv.org/content/medrxiv/early/2022/10/22/2022.10.21.22280349/F1.medium.gif) [Figure 1:](http://medrxiv.org/content/early/2022/10/22/2022.10.21.22280349/F1) Figure 1: Population structure of *S. aureus* isolates collected in this study. **A:** A maximum-likelihood phylogeny was constructed with IQ-TREE, using a GTR model and ultrafast bootstrapping (1000 bootstraps; version 2.0.3)(108, 109) from an alignment of all isolates (n=86). Scale bar represents a genetic distance of 0.001. Leaves from isolates that were “cured” are white (n=65) and those that did not achieve a “cured” status are black (n=21). The tree is annotated with MLST, clonal complex, methicillin resistance status, *SCCmec* and *spa* types and the presence of PVL genes (indicated by colored blocks). Isolates from hospital associated lineages are highlighted in blue. Interactive visualization is available on Microreact (110): [https://microreact.org/project/Post-Pascoe-Saureus/0f0c26fb](https://microreact.org/project/Post-Pascoe-Saureus/0f0c26fb). The number of isolates from each (**B**) MLST clonal complexes (CC) and (**C**) methicillin resistant (MRSA) and methicillin susceptible (MSSA) lineages. The proportion of isolates that lead to a “not cured” status is shown in black. Most of our isolates (n=48/86; 55.8%) were from community-acquired lineages (MSSA or SCC*mec* types I, II, III), including CC30 which was the most common clonal complex identified in our collection (n=14/86; 16.3%) (**Figure 1A+B**). In more than three quarters (n=11/14; 78.6%,) of cases where this clonal complex was identified, the patient was deemed to have had a good outcome and a “cured” status (**Supplementary table S2**). A minority of isolates were identified from other CA-lineages, CC8 (n=6) and CC59 (n=5), CC45 (n=9), CC7 (n=8), CC15 (n=6). No isolates were sampled from two common pandemic CA-lineages, CC1 and CC80. Many isolates we identified were from well described hospital-acquired lineages (n=23/86; 26.7%), including the globally distributed CC5 lineage (n=12) (**Figure 1A+B**). Infections from this CC were often unresolved (n=6/12; 50%) and posed the highest risk of a “not cured” patient outcome. CC22 is the most common sequence type identified from clinical infections, particularly in the UK (51, 52), but was the 3rd most sampled CC in our collection (n=11/86; 12.8%). This clonal complex was implicated in a “not cured” patient outcome on only one occasion. All other CCs were represented by fewer than 5 isolates. One isolate was isolated from the livestock associated lineage, CC398 (n=1). #### Accessory genome differences in ODRI isolates ODRI that leads to a “not cured” patient outcome is a complex process, which is affected by genetic and environmental differences in the host (**Table 2**) as well as variation in the infecting bacterial population. To investigate differences in gene presence between isolates showing phenotypic variation, we further investigated genes in the accessory genome. We characterized the pangenome using PIRATE (53) with 86 isolates plus 8 reference strains (to help preserve gene nomenclature). In total, PIRATE identified 4,142 gene clusters, of which over half were characterized as core genes (present in 95% or more of the isolates; 2,150 genes; 52%). This was consistent with core genome estimates from other *S. aureus* collections (54–57). The accessory genome consisted of 1,992 genes (present in fewer than 95% of isolates), representing ∼48% of the pangenome. A large proportion of the accessory genome (78%) was present in fewer than 25% of isolates (1,57/1,992 accessory genes; **Supplementary table S3**). Core and accessory genome differences were visualized in phandango (58) (**Figure S1A**). Consistent with the clonal nature of *S. aureus*, isolates grouped by accessory genome content (using poppunk) clustered similarly to the clonal frame (**Figure S1BC**) (59). #### Distribution of known virulence and AMR genes All *S. aureus* isolates were sampled from invasive disease cases and contained many known virulence genes, with between 50 and 69 genes identified in each isolate from the VfDB database (average: 63; update March 2021). Nearly half of these putative virulence factors were present in all isolates (35 of 79; 44%), including the genes: *adsA* (phagocytosis escape), *aur* (metalloproteinase), *geh* (lipase), *hla (*α-hemolysin*), hlgAB (*γ*-*hemolysin), *hysA* (hyaluronate lysate), *icaABDR* (biofilm formation), *isdAB* (surface proteins), *lip* (lipase), *srtB* (surface protein anchor), *sspABC* (adhesion) and much of the *cap8* capsule operon (**Figure 2A**). Distribution of virulence genes between the two clinical outcome groups (“cured” and “not cured”) was also considered. The methicillin resistance gene, *mecA* was more prevalent in the “not cured” outcome group (n=2/21; 9.5%) compared to the “cured” group (n=3/65; 4.6%). Also, the presence of the *bbp* gene (bone sialprotein binding) (n=20/21; 95.2% versus n=58/65; 89.2%) and *ebpS* gene (elastin binding) (n=21/21; 100.0% versus n=60/65; 92.3%) were found to be more prevalent in the “not cured” outcome group than in the “cured” outcome group, but this was not statistically significant. (**Supplementary table S4**). ![Figure 2:](http://medrxiv.org/https://www.medrxiv.org/content/medrxiv/early/2022/10/22/2022.10.21.22280349/F2.medium.gif) [Figure 2:](http://medrxiv.org/content/early/2022/10/22/2022.10.21.22280349/F2) Figure 2: Hypervirulent lineages and clones. **A:** Maximum-likelihood phylogeny of our 86 *S. aureus* isolates above matrix of defined virulence factors identified using the VfDb (114). Summary boxplots of the number of virulence and antimicrobial resistance genes (ARGs) identified in each (**B)** clonal complex and (**C**) MSSA clone (represented by 3 or more isolates). All data points shown, bars show min and max. Hospital-associated lineages and clones highlighted in blue. The number of isolates from each MSSA clone is also shown, with the proportion of isolates from patients who did not achieve a” cured” status in black. As most isolates were MSSA, there were very few antimicrobial resistance genes (ARGs) identified through nucleotide comparisons with the AMRfinder Plus database (**Supplementary table S5**). Differences were observed between clonal complexes for the numbers of virulence and AMR determinants (**Figure 2B**). In common with other studies, we identified slightly fewer virulence genes, but more ARGs in hospital-associated lineages. On average there were 60.3 virulence genes and 4.4 ARGs in HA lineages, compared with 63.8 virulence and 4.1 ARGs in community associated lineages. Six MSSA clones were represented by three or more isolates, of which five represented more than a third of our isolates (8 of 21; 38%) from patients who experienced a “not cured” outcome (**Figure 2C**). Two clones posed a particularly high risk to patients with 40% (2 of 5) of CC5-t002-MSSA and CC59-t216-MSSA isolates developing a “not cured” patient outcome. A quarter of patients (2 of 8) infected by CC7-t091-MSSA also experienced a “not cured” outcome, while those infected by the CC22-t005-MSSA clone all recovered. #### Both “cured” and “not cured” isolates induced high mortality in a Galleria mellonella model A selection of isolates (10 “not cured” and 10 “cured”) were used to challenge *G. mellonella* larvae. Bacterial suspensions were injected into larvae (average inoculum of 2.75×106CFU/larvae; range: 1.87×106 to 4.29× 106) and incubated up to 120 hours. However, high mortality was observed in a proportion of isolates in both outcome groups (**Figure 3AB**). This may not be surprising as all isolates were from infection cases, and all were found to possess at least 12 toxin genes. Further analysis of the *Galleria* virulence results identified differences between the average mortality scores when specific putative virulence genes were present (**Figure 3C**). Differences in gene content between isolates were quantified and scored using SCOARY (60). We report minus log10 of the naïve p-value for the null hypothesis that the presence/absence of this gene is unrelated to the trait status (-log10>2 equivalent to naive p-value <0.01). We identified four gene clusters associated with increased killing during the *Galleria* infection model (**Figure 3D**). Three of which (g01257, g01221 and g01491) demonstrated more than 90% sequence similarity with a SOS response-activated pathogenicity island shared between *S. aureus* and *S. epidermidis* isolates (SACOL0900-0904) (54, 61). The fourth gene cluster (g03516) identified was identical (100% nucleotide similarity over 100% of the gene) to the *msaC* gene (SACOL1438/SAUSA300_1296). As an uncoded member of the *msaABCR* operon, this locus has an indirect role in virulence factors *aur, scp, ssp* and *spl* and contributes to increased virulence, biofilm formation and triggers the onset of bacterial persistence (62–64)(**Supplementary file S4**). ![Figure 3:](http://medrxiv.org/https://www.medrxiv.org/content/medrxiv/early/2022/10/22/2022.10.21.22280349/F3.medium.gif) [Figure 3:](http://medrxiv.org/content/early/2022/10/22/2022.10.21.22280349/F3) Figure 3: *Galleria mellonella* virulence model. Kaplan-Meier survival curve of larvae infected with *S. aureus* isolates from **A:** “not cured” outcome patients and from **B:** “cured” outcome patients. Each line represents a different *S. aureus* isolate injected into 10 larvae per isolate. Mean survival rate was calculated per isolate. Dotted line represents survival of larvae injected with PBS; green line represents high survival rate and red line indicates a high mortality rate. **C:** Mean kill curve scores from isolates that contained each virulence factor identified by VfDb. **D:** Pangenome-wide association study comparing isolates with high kill score (above 50% at 120 hrs) vs those with low kills scores using SCOARY (60). No lineage correction was used and the phylogenetically naïve minus LOG p-value reported. Three genes from previously identified pathogenicity island and *msaC* gene from the *msaABCR* operon were associated with increased killing. #### Isolate phenotypic variation may contribute to the onset of infection, but does little to influence patient outcome We tested laboratory phenotypes associated with increased virulence and compared these with patient outcome. Individually, none of the phenotypes contributed significantly to a “not cured” patient outcome (**Table 3**). Isolates were subject to antimicrobial susceptibility profiling (29 antibiotics), biofilm formation, hemolysis activity and staphyloxanthin production (**Supplementary table S6**). No isolate was considered extensively or pan drug resistant, although 4 were considered multi-drug resistant (resistant to three or more different classes of antibiotic). More than half (n=51/86; 59.3 %) of all collected isolates were unable to form a thick biofilm under laboratory conditions. Furthermore, 61.6 % (n=53/86) of all isolates produced staphyloxanthin and 39.5 % (n=34/86) showed hemolytic activity (**Table 3**). Isolates that demonstrated hemolytic activity contribute to a “cured” patient outcome (≥90% confidence; p=0.090, **Table 3**). Isolates that were able to produce a thick biofilm were often also able to produce staphyloxanthin (n=27/35; 77.1%; **Table 4**). However, poor biofilm forming isolates that produced staphyloxanthin (n=26/51; 51.0%) were detrimentally associated with a “not cured” patient outcome (n=15/26; 57.7%). View this table: [Table 3:](http://medrxiv.org/content/early/2022/10/22/2022.10.21.22280349/T3) Table 3: Phenotypic risk factors View this table: [Table 4:](http://medrxiv.org/content/early/2022/10/22/2022.10.21.22280349/T4) Table 4: Phenotype associations in combination with enhanced biofilm formation. #### Genes associated with different virulence phenotypes Differences in gene content between phenotypes were quantified and scored using SCOARY, reporting the minus log10 of the naive p-value for the null hypothesis that the presence/absence of this gene is unrelated to the trait status (-log10>2 equivalent to naive p-value <0.01; **Supplementary table S7**). We identify genes associated with six different phenotypes, including patient outcome, biofilm formation, hemolysis, multi-drug resistance, methicillin resistance and staphyloxanthin production (**Figure 4)**. An uncharacterized membrane protein (g02811) and a gene from the SA97 virulent bacteriophage (g03902) were associated with a “not cured” patient outcome (-log10>2; naïve p-value <0.01; **Figure 4A**). Genes commonly found as part of the *SCCmec* cassette (*mecA, paaZ, upgQ* and *mecRI*) were associated with methicillin resistance, and were among the gene clusters that demonstrated the strongest association with any phenotype (-log10>6; naïve p-value <2*10−7; **Figure 4B**). Several hypothetical gene clusters were associated with biofilm formation (8 genes with naïve p-value <0.01; -log10>2; **Figure 4C**). ![Figure 4:](http://medrxiv.org/https://www.medrxiv.org/content/medrxiv/early/2022/10/22/2022.10.21.22280349/F4.medium.gif) [Figure 4:](http://medrxiv.org/content/early/2022/10/22/2022.10.21.22280349/F4) Figure 4: Pangenome-wide association studies. SCOARY was used to quantify gene presence between isolates tested for six targeted virulence phenotypes (60). No lineage correction was used, and the phylogenetically naïve minus LOG p-value reported. Isolates were scored for phenotypic differences in (**A**) patient outcome; (**B**) methicillin resistance; (**C**) biofilm formation; (**D**) hemolysis; (**E**) multidrug resistance; and (**F**) staphyloxanthin production. Genes that were highly associated with a given phenotype are labelled. Multiple toxin (blue circles) and *SCCmec* cassette (red circles) genes were associated with more than one phenotype. (**G**) Individual gene phylogenies were compared to a core phylogeny for each virulence phenotype-associated gene and consistency indices calculated. The two distributions were significantly different (Mann Whitney test; U = 78998, p < 0.0001) between virulence phenotype-associated genes (0.7766 ± 0.2418) compared with the average for all core genes (0.6943 ± 0.1408). The presence of several genes (27 genes with -log10>2, lowest p-value: 0.00017), including enterotoxins A and D and the collagen adhesin *cna* were associated with hemolysis production (**Figure 4D**). Despite only four isolates that were classed as MDR, the presence of 93 genes (lowest p-value: 4.68×10−6) were associated with phenotypically measured MDR, which included several members of the *SCCmec* cassette, the erythromycin resistance-associated *ermA* and the aminoglycoside resistance determinant, *ant1* **Figure 4E**). Several enterotoxin genes were also associated with MDR, along with the biofilm-associated gene *xerC*. The chemotaxis inhibiting gene, *chp* was the only gene associated with staphyloxanthin production in our dataset (**Figure 4F**). #### Genes linked to phenotypes that influence patient outcomes are more recombinogenic Finally, we investigated the role of recombination in isolates that led to a poorer patient outcome. On average, isolates that led to a “not cured” patient outcome demonstrated the highest rates of recombination (**Figure S2**). However, genomes of isolates from patients who were successfully ‘cured’ were predicted to have included more recombination events. This is likely influenced by a small number of highly recombinogenic genomes that were predicted to have undergone extensive recombination **(Supplementary table S8**). Genes that were associated with any of our tested phenotypes demonstrated greater clonal inheritance than genes in the core genome. Meaning that phylogenies constructed from sequences of the individual genes were consistent with the phylogeny constructed from the core genome (**Figure 4G**). The mean consistency index (CI) was significantly higher (Mann Whitney test; U = 78998, p < 0.0001) among virulence phenotype-associated genes (0.7766 ± 0.2418) compared with the average for all core genes (0.6943 ± 0.1408). Taken together, the highly clonal population structure of *S. aureus* (long branches) – particularly those genes involved in virulence – and higher rates of recombination between lineages permits co-evolution of several successful hypervirulent lineages that can contribute to “not cured” patient outcomes in ODRI. ## Discussion Despite advances in many aspects of emergency and orthopedic trauma care, ODRI persists as a challenge for the treating physicians and a significant burden for the patient. The interaction between the host immune system and invading isolates is complex and is not only affected by variation in host and bacterial genetics, but also by factors such as host wellbeing and general health (e.g. obesity). In a similar way to how the host deploys several different types of host defense mechanisms, bacteria are armed with a toolbox of different virulence factors. We used a combination of phenotype testing, an *in vivo* infection model and *in silico* genome characterization to identify lineages and virulence factors that may contribute to the risk of poor patient outcomes following ODRI. Isolates from patients who experienced a “not cured” outcome did not cluster by core or accessory genomes (**Figure 1A**). The most common clonal complex associated with a poor patient outcome in our collection was CC5 (50%; **Figure 1B**). CC5 is highly recombinogenic and able to infect multiple host species (65, 66) and is often isolated from human infections, particularly skin and bone infections (55, 67). While infection is often caused by MRSA, there is growing concern for the spread of hypervirulent MSSA clones (67, 68). Nearly all the isolates we collected were sensitive to methicillin and lacked a functional *SCCmec* cassette (9 of 12 CC5 isolates). The gain and loss of *SCCmec* cassettes by multiple *S. aureus* clones continues to blur the differentiation of community-associated (CA) and hospital-associated (HA) *S. aureus* lineages (69, 70). Since the emergence and epidemic spread of USA300 (CC8) clone in the early 2000s (30, 71, 72), CA-lineages have dominated infection surveillance studies (19, 71, 72). CA-lineages are typically thought to carry larger -more cumbersome, and less resistant -SCC*mec* types I, II and III, and often carry toxin genes such as the PVL genes. While traditional HA-*S. aureus* lineages carry smaller SCC*mec* cassettes that are more difficult to disrupt, which are maintained by successive generations. Several HA-lineages rose to prominence around the world, only to have been replaced as the most common lineages in human infections in specific regions; including the rise of CC22 in the UK and Europe (18, 73–75), and replacement of HA-MRSA-ST239 by CA-MRSA-ST59 in China (19). It is has been suggested that increased virulence, coupled with fewer putative AMR genes has contributed to the success of these CA-clones/lineages (31, 76). This was the case in our collection with more virulence genes and fewer ARGs found in CA-lineages compared to HA-lineages (**Figure 2B & Table S5**). Isolates from the most common European lineage (CC22) were common among our collection of invasive isolates and were the most likely to lead to a “cured” patient outcome (**Figure 1B**). Further characterization of the specific clones in our collection identified 6 common clones, including a CC22-t005-MSSA clone that did not lead to a “not cured” patient outcome in any of the 5 patients it infected (**Figure 2C**). The five remaining clones that were identified in three of more patients were responsible for more than a third of the “not cured” patient outcomes. The CC59-t216-MSSA and CC5-t002-MSSA clones posed the highest risk, with 40% of patients infected by these clones developing a poor outcome. The CC5-t002-MRSA clone has been frequently observed, including studies from China and Iran where it exhibited extensive mupirocin resistance, MDR and contained several virulence toxins (PVL and tsst-1) (77, 78). All isolates we collected were from invasive ODRI and contained at least 50 known virulence genes (**Figure 2AB**). Infection is complex and the function of any virulence gene is context dependent and a simple sum of the number of putative virulence elements is unlikely to be a good proxy for the infective potential of an isolate. All isolates in our collection were able to colonize and cause infection, but variation in additional virulence factors will influence disease progression and host evasion. Despite relatively small sample numbers, we were able to identify putative virulence-associated genes from an *in vivo* infection model and phenotypic assays, including biofilm production, antimicrobial susceptibility (and MDR), hemolysis and staphyloxanthin production. Similar assays in *S. epidemidis* identified biofilm formation as key to establishing an ODRI through adhesion to implant surfaces (12, 43) and contributed towards a poorer clinical outcome, persistence and recurrence (8). However, this was not the case in the current study. Although 40% of the *S. aureus* isolates were able to form a strong biofilm, no clear influence on patient outcome was observed (**Table 3**). Armed with a suite of extracellular toxins, *S. aureus* can often cause more severe acute infections (79), while *S. epidermidis* is often associated with less severe, chronic infection – where strong biofilm formation on implants and dead tissue may be more relevant (80–82). *In vitro* Staphyloxanthin production leads to increased pathogenicity of *S. aureus* (83, 84), and 61% of our isolates were able to produce staphyloxanthin (**Table S6**). This level of prevalence is consistent with the observation by Post *et al*. whereby 56% of 305 isolates from implant and non-implant infections were staphyloxanthin producers (85). The chemotaxis inhibitor protein (CHIPS), encoded by *chp* was the only gene we found associated with staphyloxanthin production (**Figure 4G**). CHIPS ability to help evade host immune responses pose the potential for more severe and chronic infections (86, 87). Although it was not independently associated with “not cured” in our study population. In hypervirulent MSSA lineages and clones, increased carriage of virulence genes is balanced against limited carriage of AMR genes. We observed this in our collection, with relatively few isolates resistant to multiple antibiotics (**Figure 2B & Table S5**), and lower rates of MRSA isolates compared to other studies (7% MRSA and 10% MDR). Rates of MRSA (27% & 24%) were much higher in other studies of ODRI (10, 11). Both these previous studies involved patients with a median age above 60, and patients infected with MRSA experienced lower cure rates (57%) than for MSSA-infected patients (72%) or CoNS (82%) (10). Very high rates of methicillin resistance have been observed in *S. epidermidis* ODRI isolates, which is consistent with different modes of infection between the species (73% MRSE and 76% MDR) (8). Elderly patients are more likely to suffer from infection with resistant bacteria due to repeated hospital stays and multi-morbidity (88, 89), and may partially explain the reduced number of MRSA isolates in our collection. Although only present in a small number of our isolates (n=6; 7%), methicillin resistance was associated with a “not cured” outcome (2 of 6; 33.3%; **Table 3**). This is consistent with many other studies that have highlighted the detrimental effect of methicillin resistance on treatment success (43, 55, 90). Deep branching clades and highly structured clustering of isolates is evidence of the strong selection pressures that have driven the evolution of *S. aureus*. Distinct *S. aureus* lineages are found at markedly different prevalence globally, including hypervirulent sub-lineages (74, 91). Waves of pandemic sub-lineages have been described within CC30 - initially MSSA, before being replaced by MRSA variants (18, 19, 74). Independent acquisition of virulence determinants in these lineages was associated with the rise and fall of these lineages and horizontal gene transfer (HGT) likely has an important role in the accumulation of virulence factors. This lineage specific accumulation of virulence genes makes traditional identification of virulence factors difficult. Genome-wide association studies that can account for these lineage effects (43, 44, 90, 92), or can differentiate between the types of association, are potentially useful (93). Furthermore, the inclusion of the pangenome analyses (including mobile elements and intergenic regions) and covarying SNPs and genes (epistasis) (94, 95), combined with appropriate phenotypic validation and/or transcriptome profiling will greatly enhance understanding of the contribution of pathogen genetic variation on disease progression (49). Infection is clearly a complex process influenced by the host immune system and genotypic variation within the invading pathogen. Here, we only tested a single *S. aureus* colony from each patient. However, there can be considerable diversity among the commensal and infective isolates within a single person (96–99). Small colony variants (SCV) (100, 101), persistence (102, 103) and dormancy (104, 105) also help protect invading cells from the host immune system and evade antimicrobial treatment. While the isolates collected here have already colonized host subcutaneous tissue and established an infection, our aim was to delineate factors that contribute to poorer patient outcomes. Our results are consistent with a complex balance between virulence and colonization (32), involving core and accessory genome elements, and provide some evidence that characterization of common emerging clones and lineages may help disease outcome prediction. ## Materials and methods ### Clinical data and Staphylococcus aureus collection The clinical data and *S. aureus* collection was part of a prospective study performed between November 2011 and September 2013 at BGU Murnau in Germany (12) which was approved by the local ethical committee *Ethik-Kommission der Bayerischen Landesärztekammer* under approval number 12063 and registered with [https://clinicaltrials.gov](https://clinicaltrials.gov) with identifier [NCT02971657](http://medrxiv.org/lookup/external-ref?link_type=CLINTRIALGOV&access_num=NCT02971657&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom). All patients enrolled in the study were aged 18 or older and provided informed written consent. Patient inclusion criteria were culture-positive *S. aureus* infection involving fracture related infection (FRI) or periprosthetic joint infections (PJI). All patient details were anonymized, and *S. aureus* isolates given genome database identifiers (BIGSids), study identifiers (ARI-number) and sample laboratory identifiers (Lab-ID). Associated clinical data are summarized in **Table S1**. After an average of 23 months follow-up (FUP) patients were assessed for treatment outcome. When patients did not complete a follow-up appointment, the clinical outcome was assessed at the time of hospital discharge. “Cured” patients were free of infection, surgical and systemic antibiotic therapy had ceased with function of the affected joint or limb restored. If one or more of these parameters were negative, patients were considered to have had a “not cured” outcome (8, 12). ### Genome sequencing and assembly *S. aureus* colonies were cultured in 5 ml Tryptone Soy Broth at 37 °C with overnight shaking. DNA was extracted using the QIAamp DNA Mini Kit (Qiagen, Germany) according to manufacturer’s instructions with the addition of 1.5 μg/μL lysostaphin (Sigma-Aldrich, Buchs, Switzerland) and 2 µg/ml lysozyme (Sigma-Aldrich, Buchs, Switzerland) to facilitate cell lysis. DNA was quantified using a spectrophotometer prior to sending for sequencing by Microsynth AG (Switzerland) using an Illumina MiSeq benchtop sequencer. Sequencing libraries were prepared using Nextera XT library preparation kits (v2) and paired end 250 bp reads generated with the MiSeq run kit (v2). Short read paired-end data was assembled *de novo* with SPAdes (version 3.3.0; using the *–careful* command)(106) and assessed for quality (n=86 genomes contained < 500 contigs). Average assembled genomes were 2,771,938 bp in length (range: 2,638,312 - 2,962,075 bp) consisting of 113 contigs (range: 29 – 358 contigs) with an N50 of 43,492 bp (range: 5,729 – 109,980) (**Table S2**). ### Genome archiving, multiple genome alignments and construction of isolate genealogies An alignment of all 86 *S. aureus* isolates was constructed from concatenated gene sequences of all core genes (found in ≥95% isolates) using MAFFT (version 7)(107) on a gene-by-gene basis (size: 2,138,455 bp; **Supplementary file 1**). A maximum-likelihood phylogenies were constructed using a GTR + I + G substitution model and ultra-fast bootstrapping (1,000 bootstraps) implemented in IQ-TREE (version 2.0.3)(108, 109) and visualized on Microreact: [https://microreact.org/project/Post-Pascoe-Saureus/0f0c26fb](https://microreact.org/project/Post-Pascoe-Saureus/0f0c26fb) (**Supplementary file S3**) (110). Collection also shared on the pathogen watch website ([https://pathogen.watch/collection/7rgjyrzz3xoc-post-and-pascoe-et-al-2022-hypervirulent-mssa-from-odri](https://pathogen.watch/collection/7rgjyrzz3xoc-post-and-pascoe-et-al-2022-hypervirulent-mssa-from-odri)). ### Molecular typing and genome characterization Our publicly available BIGSdb database ([https://sheppardlab.com/resources/](https://sheppardlab.com/resources/)) includes functionality to determine multi-locus sequence type (MLST) profiles defined by the Staphylococcal pubMLST database ([https://pubmlst.org/saureus](https://pubmlst.org/saureus); accessed March 2019) (111). Staphopia (version 1.0.0) was used to define *SCCmec* types (112) and *spa* types were typed *in silico* using spaTyper v1.0 (**Supplementary table S2**) (113). ### Care and accessory genome characterization All unique genes present in at least one of our isolates (or eight reference strains) were identified by automated annotation using PROKKA (version 1.13) followed by PIRATE, a pangenomics tool that allows for orthologue gene clustering in bacteria (**Supplementary table S3; Supplementary file S3**). Genes families in PIRATE were defined using a wide range of amino acid percentage sequence identity thresholds for Markov Cluster algorithm (MCL) clustering (45, 50, 60, 70, 80, 90, 95, 98). Core genes were defined as present in 95% of the genomes and accessory genes as present in at least one isolate (**Supplementary figure S1A**). The pangenome was visualized using phandango, as a matrix of gene presence alongside a core genome phylogeny (58). Pairwise core and accessory genome distances were compared using PopPunk (version 1.1.4), which uses pairwise nucleotide k-mer comparisons to distinguish shared sequence and gene content to identify divergence of the accessory genome in relation to the core genome. A two-component Gaussian mixture model was used to construct a network to define clusters and visualized with microreact (**Supplementary figure S1BC**) (59, 110). ### Identification of known virulence traits and antimicrobial resistance genes The presence of putative virulence factors and antimicrobial resistance genes were identified from assembled genomes using ABRICATE (v0.3) (59). Putative virulence genes were detected through comparison with reference nucleotide sequences in the VfDb (default settings: ≥70% identity over ≥50% of the gene; **Supplementary table S4**) (114). Antimicrobial resistance genes were identified using the NCBI AMRfinder Plus (115) and CARD (116) databases (19th April, 2020 update). Results for both AMR databases were similar, with the CARD database identifying additional AMR associated genes. We report the results from the curated AMRfinder Plus database (**Supplementary table S5**). ### Phenotype testing ***Antibiotic susceptibility testing:*** Minimum inhibitory concentrations (MICs) for 29 antibiotics were determined using a Vitek2 machine (bioMérieux Vitek Inc., USA) as described previously (8, 12). Susceptibility breakpoints were defined according to the definitions of the European Committee of Antimicrobial Susceptibility Testing (EUCAST) and isolates resistant to 3 or more antimicrobial classes were defined as multidrug resistant (MDR)(118). The Vitek2 results for oxacillin and cefoxitin were used as a proxy for methicillin-resistance. ***Staphyloxanthin production:*** Staphyloxanthin production indicated by a yellow-orange pigmentation of *S. aureus* colonies as well as ***Haemolysin activity:*** Haemolysis activity of *S. aureus* strains was evaluated as previously described (119). ***Biofilm formation***: was also assayed as described previously (8, 12, 85, 120). Association amongst and between the clinical parameters, bacterial phenotypes, clades and presence/absence of genes were analyzed statistically using Chi-square test. Statistical analyses were performed using SPSS (Version 23, IBM, USA) or GraphPad prism 6 (GraphPad Software, Inc.; **Supplementary table S6**). ***In vivo virulence and survival assay in Galleria mellonella larvae:*** The invertebrate *G. mellonella* infection model was used to study the virulence of the *S. aureus* strains as previously described (121, 122). *G. mellonella* were obtained at pre-larval stage (Entomos AG, Zurich Switzerland). Larvae were grown at 30 °C in the dark and groups of ten larvae in the final instar larval stage weighing 200– 400Lmg were used in all assays. A bacterial suspension of 106 CFU/ml was prepared. Quantitative culture of a sample from each bacterial suspension was performed immediately after preparation by ten-fold serial dilution and plating on Tryptone Soy Agar plates to check the actual total viable count of the prepared suspension. Bacterial inoculates (10 µl) were injected into the last left pro-leg into the hemocoel of the last-instar larvae (200-400 mg). After injection, larvae were incubated at 37 °C in the dark. Larvae were assessed daily for survival up to 5 days post-injection and were evaluated according to survival, being scored as dead when they displayed no movement in response to touch. Controls included a group of larvae that did not receive any injection and a group of larvae inoculated with sterile phosphate-buffered saline (PBS). Experiments consisted of 10 larvae per bacterial strain, which was repeated in three separate experiments. For *G. mellonella* survival analysis, larvae mean survival curves were plotted using the Kaplan-Meier method (GraphPad Prism 6, USA; **Supplementary table S6**). ### Pan-genome-wide association studies of phenotype variation Pangenome-wide differences in gene presence were quantified using the genome-wide association software, SCOARY (version 1.6.14)(60). With only limited numbers of samples in our collection, we report phylogenetically naïve differences in gene presence between clinical and laboratory symptoms and phenotypes (**Supplementary table S7**). ### Recombination and allelic consistency The number of polymorphisms introduced by mutation and recombination in the core genome were inferred using Gubbins (version 2.4.1)(123) for each isolate (per branch; **Supplementary table S8**).The consistency of the phylogenetic tree to patterns of variation in sequence alignments for each gene of interest was calculated as before (124, 125). Consistency indices for each single-gene alignment of 130 virulence-associated genes to a phylogeny constructed using an alignment of 2,150 core genes shared by 94 isolates, were calculated using the CI function of the R Phangorn package (126). ## Supporting information Figure S1 [[supplements/280349_file11.pdf]](pending:yes) Figure S2 [[supplements/280349_file12.pdf]](pending:yes) Supplementary table 1 [[supplements/280349_file13.xlsx]](pending:yes) Supplementary table 2 [[supplements/280349_file14.xlsx]](pending:yes) Supplementary table 3 [[supplements/280349_file15.xlsx]](pending:yes) Supplementary table 4 [[supplements/280349_file16.xlsx]](pending:yes) Supplementary table 5 [[supplements/280349_file17.xlsx]](pending:yes) Supplementary table 6 [[supplements/280349_file18.xlsx]](pending:yes) Supplementary table 7 [[supplements/280349_file19.xlsx]](pending:yes) Supplementary table 8 [[supplements/280349_file20.xlsx]](pending:yes) ## Data Availability Illumina short read sequence data are archived on the Sequence Read Archive. All assembled genomes are shared on figshare (doi.org: 10.6084/m9.figshare.7926866) and are available on our public Staphylococcal Bacterial Isolate Genome Sequence Database (BIGSdb): https://sheppardlab.com/resources/. Collection also shared on the pathogen watch website (https://pathogen.watch/collection/7rgjyrzz3xoc-post-and-pascoe-et-al-2022-hypervirulent-mssa-from-odri). Isolate genealogy with associated meta-data is visualized on microreact [110]: https://microreact.org/project/Post-Pascoe-Saureus/0f0c26fb. ## Data availability Illumina short read sequence data are archived on the Sequence Read Archive associated with BioProject accession PRJNA529795. All assembled genomes are shared on figshare (doi: 10.6084/m9.figshare.7926866) and are available on our public Staphylococcal Bacterial Isolate Genome Sequence Database (BIGSdb): [https://sheppardlab.com/resources/](https://sheppardlab.com/resources/). Isolate genealogy with associated meta-data is visualized on microreact (110): [https://microreact.org/project/Post-Pascoe-Saureus/0f0c26fb](https://microreact.org/project/Post-Pascoe-Saureus/0f0c26fb). All high-performance computing was performed on MRC CLIMB (127) in a bioconda environment (128). ## Contributions Conceptualization: VP, BP, SKS, and TFM; Isolate collection: VP, CE, JF, MM, and TFM; Data collection and curation: VP, BP, MDH and SKS; Formal Analysis: VP, BP, EM, JKC; Visualization: VP, BP, EM, JKC; Writing – original draft: VP, BP, SKS and TFM; Writing – Review & Editing: VP, BP, SKS and TFM; Funding Acquisition: VP, RGR and TFM. ## Figures and Table legends **Table 1:** Collection overview. **Table 2:** Patient risk factors. **Table 3:** Infection isolate prognostic phenotypes. **Table 4:** Phenotype associations in combination with enhanced biofilm formation. **Supplementary material (**figshare: [https://doi.org/10.6084/m9.figshare.7926866](https://doi.org/10.6084/m9.figshare.7926866)**)** **Table S1:** Demographic data collected for each isolate. **Table S2:** Genome quality control and extended typing results. **Table S3:** Summary of core and accessory genome characterization with PIRATE. **Table S4:** Summary of genes identified by screening known virulence genes with the VfDB. **Table S5:** Summary of genes identified by screening antibiotic resistance genes with the AMRfinder (NCBI) and CARD databases. **Table S6:** Summary of virulence phenotype data. **Table S7:** Summary of genes associated with virulence phenotypes (SCOARY). **Table S8:** Per isolate recombination statistics using Gubbins. **Figure S1:** Accessory genome characterization. **Figure S2**: Recombination analysis. **Supplementary file 1:** Contigs **Supplementary file 2:** Alignment. **Supplementary file 3:** Phylogeny (.nwk) of 86 collected isolates. **Supplementary file 4**: PIRATE pangenome gene families. ## Acknowledgements This work was funded by AO Trauma as part of the Clinical Priority Program, Bone Infection. All high-performance computing was performed on MRC CLIMB, funded by the Medical Research Council (MR/L015080/1 & MR/T030062/1). This publication made use of the PubMLST website ([http://pubmlst.org/](http://pubmlst.org/)) developed by Keith Jolley and Martin Maiden (129) and sited at the University of Oxford. The development of that website was funded by the Wellcome Trust. The funders had no role in study design, data collection and interpretation, or the decision to submit the work for publication. All authors report no conflicts of interest relevant to this article. * Received October 21, 2022. * Revision received October 21, 2022. * Accepted October 22, 2022. * © 2022, Posted by Cold Spring Harbor Laboratory This pre-print is available under a Creative Commons License (Attribution 4.0 International), CC BY 4.0, as described at [http://creativecommons.org/licenses/by/4.0/](http://creativecommons.org/licenses/by/4.0/) ## References 1. 1.Tschudin-Sutter S, Frei R, Dangel M, Jakob M, Balmelli C, Schaefer DJ, Weisser M, Elzi L, Battegay M, Widmer AF. 2016. Validation of a treatment algorithm for orthopaedic implant-related infections with device-retention— results from a prospective observational cohort study. Clin Microbiol Infect 22:457.e1-457.e9. 2. 2.Zimmerli W. 2014. Clinical presentation and treatment of orthopaedic implant-associated infection. J Intern Med. 3. 3.Moriarty TF, Kuehl R, Coenye T, Metsemakers W-JJ, Morgenstern M, Schwarz EM, Riool M, Zaat SAJJ, Khana N, Kates SL, Geoff Richards R, Richards RG. 2016. Orthopaedic device-related infection: Current and future interventions for improved prevention and treatment. EFORT Open Rev 1:89–99. [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MzoiZW9yIjtzOjU6InJlc2lkIjtzOjY6IjEvNC84OSI7czo0OiJhdG9tIjtzOjUwOiIvbWVkcnhpdi9lYXJseS8yMDIyLzEwLzIyLzIwMjIuMTAuMjEuMjIyODAzNDkuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 4. 4.Li B, Webster TJ. 2018. Bacteria antibiotic resistance: New challenges and opportunities for implant-associated orthopedic infections. J Orthop Res 36:22–32. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1002/jor.23656&link_type=DOI) 5. 5.Patzakis MJ, Wilkins J. 1989. Factors Influencing Infection Rate in Open Fracture Wounds. Clin Orthop Relat Res 243. 6. 6.Boxma H, Broekhuizen T, Patka P, Oosting H. 1996. Randomised controlled trial of single-dose antibiotic prophylaxis in surgical treatment of closed fractures: The Dutch Trauma Trial. Lancet 347:1133–1137. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1016/S0140-6736(96)90606-6&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=8609746&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=A1996UG75500008&link_type=ISI) 7. 7.Spiegl U, Friederichs J, Pätzold R, Militz M, Josten C, Bühren V. 2013. Risk factors for failed two-stage procedure after chronic posttraumatic periprosthetic hip infections. Arch Orthop Trauma Surg 133:421–428. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1007/s00402-012-1673-6&link_type=DOI) 8. 8.Morgenstern M, Post V, Erichsen C, Hungerer S, Bühren V, Militz M, Richards RG, Moriarty TF. 2016. Biofilm formation increases treatment failure in Staphylococcus epidermidis device-related osteomyelitis of the lower extremity in human patients. J Orthop Res 34:1905–1913. 9. 9.Guo G, Wang J, You Y, Tan J, Shen H. 2017. Distribution characteristics of Staphylococcus spp. in different phases of periprosthetic joint infection: A review. Exp Ther Med 13:2599. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.3892/etm.2017.4300&link_type=DOI) 10. 10.Teterycz D, Ferry T, Lew D, Stern R, Assal M, Hoffmeyer P, Bernard L, Uçkay 2010. Outcome of orthopedic implant infections due to different staphylococci. Int J Infect Dis 14. 11. 11.Kilgus DJ, Howe DJ, Strang A. 2002. Results of Periprosthetic Hip and Knee. Infections Caused by Resistant Bacteria. Clin Orthop Relat Res 404:116–124. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1097/00003086-200211000-00021&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=12439249&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) 12. 12.Post V, Harris LG, Morgenstern M, Mageiros L, Hitchings MD, Méric G, Pascoe B, Sheppard SK, Richards RG, Moriarty TF. 2017. Comparative Genomics Study of Staphylococcus epidermidis Isolates from Orthopedic-Device-Related Infections Correlated with Patient Outcome. J Clin Microbiol 55:3089–3103. [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MzoiamNtIjtzOjU6InJlc2lkIjtzOjEwOiI1NS8xMC8zMDg5IjtzOjQ6ImF0b20iO3M6NTA6Ii9tZWRyeGl2L2Vhcmx5LzIwMjIvMTAvMjIvMjAyMi4xMC4yMS4yMjI4MDM0OS5hdG9tIjt9czo4OiJmcmFnbWVudCI7czowOiIiO30=) 13. 13.Arciola CR, An YH, Campoccia D, Donati ME, Montanaro L. 2005. Etiology of implant orthopedic infections: A survey on 1027 clinical isolates. Int J Artif Organs 28:1091–1100. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1177/039139880502801106&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=16353115&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000233814800006&link_type=ISI) 14. 14.Campoccia D, Montanaro L, Arciola CR. 2006. The significance of infection related to orthopedic devices and issues of antibiotic resistance. Biomaterials 27:2331–2339. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1016/j.biomaterials.2005.11.044&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=16364434&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000235631500001&link_type=ISI) 15. 15.Montanaro L, Ravaioli S, Ruppitsch W, Campoccia D, Pietrocola G, Visai L, Speziale P, Allerberger F, Arciola CR. 2016. Molecular characterization of a prevalent ribocluster of methicillin-sensitive Staphylococcus aureus from Orthopedic implant infections. Correspondence with MLST CC30. Front Cell Infect Microbiol 6:8. 16. 16.Otto M. 2012. Methicillin-resistant Staphylococcus aureus infection is associated with increased mortality. [http://dx.doi.org/102217/fmb11156](http://dx.doi.org/102217/fmb11156) 7:189–191. 17. 17.Zhuang H, Zhu F, Lan P, Ji S, Sun L, Chen Y, Wang Z, Jiang S, Zhang L, Zhu Y, Jiang Y, Chen Y, Yu Y. 2021. A random forest model based on core genome allelic profiles of mrsa for penicillin plus potassium clavulanate susceptibility prediction. Microb Genomics 7:000610. 18. 18.Chambers HF, DeLeo FR. 2009. Waves of Resistance: Staphylococcus aureus in the Antibiotic Era. Nat Rev Microbiol 7:629. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1038/nrmicro2200&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=19680247&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000268942500010&link_type=ISI) 19. 19.Chen H, Yin Y, van Dorp L, Shaw LP, Gao H, Acman M, Yuan J, Chen F, Sun S, Wang X, Li S, Zhang Y, Farrer RA, Wang H, Balloux F. 2021. Drivers of methicillin-resistant Staphylococcus aureus (MRSA) lineage replacement in China. Genome Med 13:1–14. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1186/S13073-021-01003-9/METRICS&link_type=DOI) 20. 20.Harkins CP, Pichon B, Doumith M, Parkhill J, Westh H, Tomasz A, de Lencastre H, Bentley SD, Kearns AM, Holden MTG. 2017. Methicillin-resistant Staphylococcus aureus emerged long before the introduction of methicillin into clinical practice. Genome Biol 18:1–11. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1186/s13059-016-1145-3&link_type=DOI) 21. 21.Enright MC, Day NPJ, Davies CE, Peacock SJ, Spratt BG. 2000. Multilocus sequence typing for characterization of methicillin-resistant and methicillinsusceptible clones of Staphylococcus aureus. J Clin Microbiol 38:1008–1015. [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MzoiamNtIjtzOjU6InJlc2lkIjtzOjk6IjM4LzMvMTAwOCI7czo0OiJhdG9tIjtzOjUwOiIvbWVkcnhpdi9lYXJseS8yMDIyLzEwLzIyLzIwMjIuMTAuMjEuMjIyODAzNDkuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 22. 22.Feil EJ, Cooper JE, Grundmann H, Robinson DA, Enright MC, Berendt T, Peacock SJ, Smith JM, Murphy M, Spratt BG, Moore CE, Day NPJ. 2003. How clonal is Staphylococcus aureus? J Bacteriol 185:3307–3316. [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MjoiamIiO3M6NToicmVzaWQiO3M6MTE6IjE4NS8xMS8zMzA3IjtzOjQ6ImF0b20iO3M6NTA6Ii9tZWRyeGl2L2Vhcmx5LzIwMjIvMTAvMjIvMjAyMi4xMC4yMS4yMjI4MDM0OS5hdG9tIjt9czo4OiJmcmFnbWVudCI7czowOiIiO30=) 23. 23.Urwin R, Maiden MCJ. 2003. Multi-locus sequence typing: a tool for global epidemiology. Trends Microbiol 11:479–487. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1016/j.tim.2003.08.006&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=14557031&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000186206900007&link_type=ISI) 24. 24.Tenover FC, Goering R V. 2009. Methicillin-resistant Staphylococcus aureus strain USA300: origin and epidemiology. J Antimicrob Chemother 64:441–446. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1093/jac/dkp241&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=19608582&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000273181400002&link_type=ISI) 25. 25.Ito T, Hiramatsu K, Oliveira DC, De Lencastre H, Zhang K, Westh H, O’Brien F, Giffard PM, Coleman D, Tenover FC, Boyle-Vavra S, Skov RL, Enright MC, Kreiswirth B, Kwan SK, Grundmann H, Laurent F, Sollid JE, Kearns AM, Goering R, John JF, Daum R, Soderquist B. 2009. Classification of staphylococcal cassette chromosome mec (SCCmec): Guidelines for reporting novel SCCmec elements. Antimicrob Agents Chemother 53:4961–4967. [FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiRlVMTCI7czoxMToiam91cm5hbENvZGUiO3M6MzoiYWFjIjtzOjU6InJlc2lkIjtzOjEwOiI1My8xMi80OTYxIjtzOjQ6ImF0b20iO3M6NTA6Ii9tZWRyeGl2L2Vhcmx5LzIwMjIvMTAvMjIvMjAyMi4xMC4yMS4yMjI4MDM0OS5hdG9tIjt9czo4OiJmcmFnbWVudCI7czowOiIiO30=) 26. 26.Strommenger B, Braulke C, Heuck D, Schmidt C, Pasemann B, Nübel U, Witte W. 2008. spa Typing of Staphylococcus aureus as a Frontline Tool in Epidemiological Typing. J Clin Microbiol 46:574. [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MzoiamNtIjtzOjU6InJlc2lkIjtzOjg6IjQ2LzIvNTc0IjtzOjQ6ImF0b20iO3M6NTA6Ii9tZWRyeGl2L2Vhcmx5LzIwMjIvMTAvMjIvMjAyMi4xMC4yMS4yMjI4MDM0OS5hdG9tIjt9czo4OiJmcmFnbWVudCI7czowOiIiO30=) 27. 27.Koeck M, Como-Sabetti K, Boxrud D, Dobbins G, Glennen A, Anacker M, Jawahir S, See I, Lynfield R. 2019. Burdens of Invasive Methicillin-Susceptible and Methicillin-Resistant Staphylococcus aureus Disease, Minnesota, USA. Emerg Infect Dis 25:171. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.3201/eid2501.181146&link_type=DOI) 28. 28.David MZ, Boyle-Vavra S, Zychowski DL, Daum RS. 2011. Methicillin-Susceptible Staphylococcus aureus as a Predominantly Healthcare-Associated Pathogen: A Possible Reversal of Roles? PLoS One 6. 29. 29.Otto M. 2013. Community-associated MRSA: What makes them special? Int J Med Microbiol 303:324–330. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1016/j.ijmm.2013.02.007&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=23517691&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) 30. 30.Uhlemann AC, Otto M, Lowy FD, DeLeo FR. 2014. Evolution of community- and healthcare-associated methicillin-resistant Staphylococcus aureus. Infect Genet Evol 21:563–574. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1016/j.meegid.2013.04.030&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=23648426&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) 31. 31.Laabei M, Uhlemann AC, Lowy FD, Austin ED, Yokoyama M, Ouadi K, Feil E, Thorpe HA, Williams B, Perkins M, Peacock SJ, Clarke SR, Dordel J, Holden M, Votintseva AA, Bowden R, Crook DW, Young BC, Wilson DJ, Recker M, Massey RC. 2015. Evolutionary Trade-Offs Underlie the Multi-faceted Virulence of Staphylococcus aureus. PLOS Biol 13:e1002229. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1371/journal.pbio.1002229&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=26331877&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) 32. 32.Sheppard SK. 2022. Strain wars and the evolution of opportunistic pathogens. Curr Opin Microbiol 67. 33. 33.Otto M. 2012. MRSA virulence and spread. Cell Microbiol 14:1513–1521. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1111/j.1462-5822.2012.01832.x&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=22747834&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) 34. 34.Otto M. 2010. Basis of virulence in community-associated methicillin-resistant Staphylococcus aureus. Annu Rev Microbiol 64:143–162. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1146/annurev.micro.112408.134309&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=20825344&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000284030600008&link_type=ISI) 35. 35.Archer GL. 1998. Staphylococcus aureus: a well-armed pathogen. Clin Infect Dis 26:1179–1181. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1086/520289&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=9597249&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000073518600029&link_type=ISI) 36. 36.Arciola CR, Visai L, Testoni F, Arciola S, Campoccia D, Speziale P, Montanaro L. 2011. Concise survey of Staphylococcus aureus virulence factors that promote adhesion and damage to peri-implant tissues. Int J Artif Organs 34:771–780. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.5301/ijao.5000046&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=22094556&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) 37. 37.Lina G, Quaglia A, Reverdy ME, Leclercq R, Vandenesch F, Etienne J. 1999. Distribution of genes encoding resistance to macrolides, lincosamides, and streptogramins among staphylococci. Antimicrob Agents Chemother 43:1062–1066. [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MzoiYWFjIjtzOjU6InJlc2lkIjtzOjk6IjQzLzUvMTA2MiI7czo0OiJhdG9tIjtzOjUwOiIvbWVkcnhpdi9lYXJseS8yMDIyLzEwLzIyLzIwMjIuMTAuMjEuMjIyODAzNDkuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 38. 38.Mehrotra M, Wang G, Johnson WM. 2000. Multiplex PCR for detection of genes for Staphylococcus aureus enterotoxins, exfoliative toxins, toxic shock syndrome toxin 1, and methicillin resistance. J Clin Microbiol 38:1032–1035. [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MzoiamNtIjtzOjU6InJlc2lkIjtzOjk6IjM4LzMvMTAzMiI7czo0OiJhdG9tIjtzOjUwOiIvbWVkcnhpdi9lYXJseS8yMDIyLzEwLzIyLzIwMjIuMTAuMjEuMjIyODAzNDkuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 39. 39.Kobras CM, Fenton AK, Sheppard SK. 2021. Next-generation microbiology: from comparative genomics to gene function. Genome Biol 2021 221 22:1–16. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1186/s13059-020-02238-2&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=33402206&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) 40. 40.Sheppard SK, Guttman DS, Fitzgerald JR. 2018. Population genomics of bacterial host adaptation. Nat Rev Genet 2018 199 19:549–565. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1038/s41576-018-0032-z&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=29973680&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) 41. 41.Méric G, Yahara K, Mageiros L, Pascoe B, Maiden MCJ, Jolley KA, Sheppard SK. 2014. A reference pan-genome approach to comparative bacterial genomics: Identification of novel epidemiological markers in pathogenic Campylobacter. PLoS One 9. 42. 42.Recker M, Laabei M, Toleman MS, Reuter S, Saunderson RB, Blane B, Török ME, Ouadi K, Stevens E, Yokoyama M, Steventon J, Thompson L, Milne G, Bayliss S, Bacon L, Peacock SJ, Massey RC. 2017. Clonal differences in Staphylococcus aureus bacteraemia-associated mortality. Nat Microbiol 2:1381–1388. 43. 43.Meric G, Mageiros L, Pensar J, Laabei M, Yahara K, Pascoe B, Kittiwan N, Tadee P, Post V, Lamble S, Bowden R, Bray JE, Morgenstern M, Jolley KA, Maiden MCJ, Feil EJ, Didelot X, Miragaia M, de Lencastre H, Moriarty TF, Rohde H, Massey RC, Mack D, Corander J, Sheppard SK. 2018. Disease-associated genotypes of the commensal skin bacterium Staphylococcus epidermidis. Nat Commun 9. 44. 44.Laabei M, Recker M, Rudkin JK, Aldeljawi M, Gulay Z, Sloan TJ, Williams P, Endres JL, Bayles KW, Fey PD, Yajjala VK, Widhelm T, Hawkins E, Lewis K, Parfett S, Scowen L, Peacock SJ, Holden M, Wilson D, Read TD, Van Den Elsen J, Priest NK, Feil EJ, Hurst LD, Josefsson E, Massey RC. 2014. Predicting the virulence of MRSA from its genome sequence. Genome Res 24:839–849. [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NjoiZ2Vub21lIjtzOjU6InJlc2lkIjtzOjg6IjI0LzUvODM5IjtzOjQ6ImF0b20iO3M6NTA6Ii9tZWRyeGl2L2Vhcmx5LzIwMjIvMTAvMjIvMjAyMi4xMC4yMS4yMjI4MDM0OS5hdG9tIjt9czo4OiJmcmFnbWVudCI7czowOiIiO30=) 45. 45.Gordon NC, Price JR, Cole K, Everitt R, Morgan M, Finney J, Kearns AM, Pichon B, Young B, Wilson DJ, Llewelyn MJ, Paul J, Peto TEA, Crook DW, Walker AS, Golubchik T. 2014. Prediction of staphylococcus aureus antimicrobial resistance by whole-genome sequencing. J Clin Microbiol 52:1182–1191. [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MzoiamNtIjtzOjU6InJlc2lkIjtzOjk6IjUyLzQvMTE4MiI7czo0OiJhdG9tIjtzOjUwOiIvbWVkcnhpdi9lYXJseS8yMDIyLzEwLzIyLzIwMjIuMTAuMjEuMjIyODAzNDkuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 46. 46.Berthenet E, Yahara K, Thorell K, Pascoe B, Meric G, Mikhail JM, Engstrand L, Enroth H, Burette A, Megraud F, Varon C, Atherton JC, Smith S, Wilkinson TS, Hitchings MD, Falush D, Sheppard SK. 2018. A GWAS on Helicobacter pylori strains points to genetic variants associated with gastric cancer risk. BMC Biol [https://doi.org/10.1186/s12915-018-0550-3](https://doi.org/10.1186/s12915-018-0550-3). 47. 47.Hwang W, Yong JH, Min KB, Lee KM, Pascoe B, Sheppard SK, Yoon SS. 2021. Genome-wide association study of signature genetic alterations among pseudomonas aeruginosa cystic fibrosis isolates. PLoS Pathog 17. 48. 48.Peters S, Pascoe B, Wu Z, Bayliss SC, Zeng X, Edwinson A, Veerabadhran-Gurunathan S, Jawahir S, Calland JK, Mourkas E, Patel R, Wiens T, Decuir M, Boxrud D, Smith K, Parker CT, Farrugia G, Zhang Q, Sheppard SK, Grover M. 2021. Campylobacter jejuni genotypes are associated with post-infection irritable bowel syndrome in humans. Commun Biol 4. 49. 49.Sassi M, Bronsard J, Pascreau G, Emily M, Donnio P-Y, Revest M, Felden B, Wirth T, Augagneur Y. 2022. Forecasting Staphylococcus aureus Infections Using Genome-Wide Association Studies, Machine Learning, and Transcriptomic Approaches. mSystems [https://doi.org/10.1128/MSYSTEMS.00378-22](https://doi.org/10.1128/MSYSTEMS.00378-22). 50. 50.Aanensen DM, Feil EJ, Holden MTG, Dordel J, Yeats CA, Fedosejev A, Goater R, Castillo-Ramírez S, Corander J, Colijn C, Chlebowicz MA, Schouls L, Heck M, Pluister G, Ruimy R, Kahlmeter G, Åhman J, Matuschek E, Friedrich AW, Parkhill J, Bentley SD, Spratt BG, Grundmannj H, Krziwanek K, Stumvoll S, Koller W, Denis O, Struelens M, Nashev D, Budimir A, Kalenic S, Pieridou-Bagatzouni D, Jakubu V, Zemlickova H, Westh H, Larsen AR, Skov R, Laurent F, Ettienne J, Strommenger B, Witte W, Vourli S, Vatopoulos A, Vainio A, Vuopio-Varkila J, Fuzi M, Ungvári E, Murchan S, Rossney A, Miklasevics E, Balode A, Haraldsson G, Kristinsson KG, Monaco M, Pantosti A, Borg M, Van Santen-Verheuvel M, Huijsdens X, Marstein L, Jacobsen T, Simonsen GS, Airesde-Sousa M, De Lencastre H, Luczak-Kadlubowska A, Hryniewicz W, Straut M, Codita I, Perez-Vazquez M, Iglesias JO, Spik VC, Mueller-Premru M, Haeggman S, Olsson-Liljequist B, Ellington M, Kearns A. 2016. Whole-genome sequencing for routine pathogen surveillance in public health: A population snapshot of invasive Staphylococcus aureus in Europe. MBio 7. 51. 51.Garvey MI, Bradley CW, Holden KL, Oppenheim B. 2017. Outbreak of clonal complex 22 Panton-Valentine leucocidin-positive methicillin-resistant Staphylococcus aureus. J Infect Prev 18:224–230. 52. 52.de Vos AS, de Vlas SJ, Lindsay JA, Kretzschmar MEE, Knight GM. 2021. Understanding MRSA clonal competition within a UK hospital; the possible importance of density dependence. Epidemics 37:100511. 53. 53.Sc B, Ha T, Nm C, Sk S, Ej F. 2019. PIRATE: A fast and scalable pangenomics toolbox for clustering diverged orthologues in bacteria. Gigascience 8. 54. 54.Méric G, Miragaia M, De Been M, Yahara K, Pascoe B, Mageiros L, Mikhail J, Harris LG, Wilkinson TS, Rolo J, Lamble S, Bray JE, Jolley KA, Hanage WP, Bowden R, Maiden MCJ, Mack D, De Lencastre H, Feil EJ, Corander J, Sheppard SK. 2015. Ecological overlap and horizontal gene transfer in Staphylococcus aureus and Staphylococcus epidermidis. Genome Biol Evol 7. 55. 55.Manara S, Pasolli E, Dolce D, Ravenni N, Campana S, Armanini F, Asnicar F, Mengoni A, Galli L, Montagnani C, Venturini E, Rota-Stabelli O, Grandi G, Taccetti G, Segata N. 2018. Whole-genome epidemiology, characterisation, and phylogenetic reconstruction of Staphylococcus aureus strains in a paediatric hospital. Genome Med 2018 101 10:1–19. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1186/s13073-017-0512-3&link_type=DOI) 56. 56.Moller AG, Robert A. Petit I, Read TD. 2022. Species-Scale Genomic Analysis of Staphylococcus aureus Genes Influencing Phage Host Range and Their Relationships to Virulence and Antibiotic Resistance Genes. mSystems 7. 57. 57.Suzuki H, Lefébure T, Bitar PP, Stanhope MJ. 2012. Comparative genomic analysis of the genus Staphylococcus including Staphylococcus aureus and its newly described sister species Staphylococcus simiae. BMC Genomics 13:38. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1186/1471-2164-13-38&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=22272658&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) 58. 58.Hadfield J, Croucher NJ, Goater RJ, Abudahab K, Aanensen DM, Harris SR. 2018. Phandango: An interactive viewer for bacterial population genomics. Bioinformatics [https://doi.org/10.1093/bioinformatics/btx610](https://doi.org/10.1093/bioinformatics/btx610). 59. 59.Lees JA, Harris SR, Tonkin-Hill G, Gladstone RA, Lo SW, Weiser JN, Corander J, Bentley SD, Croucher NJ. 2019. Fast and flexible bacterial genomic epidemiology with PopPUNK. Genome Res 29:304–316. [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NjoiZ2Vub21lIjtzOjU6InJlc2lkIjtzOjg6IjI5LzIvMzA0IjtzOjQ6ImF0b20iO3M6NTA6Ii9tZWRyeGl2L2Vhcmx5LzIwMjIvMTAvMjIvMjAyMi4xMC4yMS4yMjI4MDM0OS5hdG9tIjt9czo4OiJmcmFnbWVudCI7czowOiIiO30=) 60. 60.Brynildsrud O, Bohlin J, Scheffer L, Eldholm V. 2016. Rapid scoring of genes in microbial pan-genome-wide association studies with Scoary. Genome Biol 17:238. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1186/s13059-016-1108-8&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=27887642&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) 61. 61.Anderson KL, Roberts C, Disz T, Vonstein V, Hwang K, Overbeek R, Olson PD, Projan SJ, Dunman PM. 2006. Characterization of the Staphylococcus aureus Heat Shock, Cold Shock, Stringent, and SOS Responses and Their Effects on Log-Phase mRNA Turnover. J Bacteriol 188:6739. [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MjoiamIiO3M6NToicmVzaWQiO3M6MTE6IjE4OC8xOS82NzM5IjtzOjQ6ImF0b20iO3M6NTA6Ii9tZWRyeGl2L2Vhcmx5LzIwMjIvMTAvMjIvMjAyMi4xMC4yMS4yMjI4MDM0OS5hdG9tIjt9czo4OiJmcmFnbWVudCI7czowOiIiO30=) 62. 62.C BG, Sahukhal GS, Elasri MO. 2022. Delineating the Role of the msaABCR Operon in Staphylococcal Overflow Metabolism. Front Microbiol 13:914512. 63. 63.Sahukhal GS, Elasri MO. 2014. Identification and characterization of an operon, msaABCR, that controls virulence and biofilm development in Staphylococcus aureus. BMC Microbiol 14:154. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1186/1471-2180-14-154&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=24915884&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) 64. 64.Pandey S, Sahukhal GS, Elasri MO. 2021. The msaABCR Operon Regulates Persister Formation by Modulating Energy Metabolism in Staphylococcus aureus. Front Microbiol 12:881. 65. 65.Murray S, Pascoe B, Méric G, Mageiros L, Yahara K, Hitchings MD, Friedmann Y, Wilkinson TS, Gormley FJ, Mack D, Bray JE, Lamble S, Bowden R, Jolley KA, Maiden MCJ, Wendlandt S, Schwarz S, Corander J, Fitzgerald JR, Sheppard SK. 2017. Recombination-Mediated Host Adaptation by Avian Staphylococcus aureus. Genome Biol Evol 9:830–842. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1093/gbe/evx037&link_type=DOI) 66. 66.Lowder B V., Guinane CM, Zakour NLB, Weinert LA, Conway-Morris A, Cartwright RA, Simpson AJ, Rambaut A, Nübel U, Fitzgerald JR. 2009. Recent human-to-poultry host jump, adaptation, and pandemic spread of Staphylococcus aureus. Proc Natl Acad Sci U S A 106:19545–19550. [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NDoicG5hcyI7czo1OiJyZXNpZCI7czoxMjoiMTA2LzQ2LzE5NTQ1IjtzOjQ6ImF0b20iO3M6NTA6Ii9tZWRyeGl2L2Vhcmx5LzIwMjIvMTAvMjIvMjAyMi4xMC4yMS4yMjI4MDM0OS5hdG9tIjt9czo4OiJmcmFnbWVudCI7czowOiIiO30=) 67. 67.Jian Y, Zhao L, Zhao N, Lv HY, Liu Y, He L, Liu Q, Li M. 2021. Increasing prevalence of hypervirulent ST5 methicillin susceptible Staphylococcus aureus subtype poses a serious clinical threat. Emerg Microbes Infect 10:109. 68. 68.Nübel U, Roumagnac P, Feldkamp M, Song JH, Ko KS, Huang YC, Coombs G, Ip M, Westh H, Skov R, Struelens MJ, Goering R V., Strommenger B, Weller A, Witte W, Achtman M. 2008. Frequent emergence and limited geographic dispersal of methicillin-resistant Staphylococcus aureus. Proc Natl Acad Sci U S A 105:14130–14135. [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NDoicG5hcyI7czo1OiJyZXNpZCI7czoxMjoiMTA1LzM3LzE0MTMwIjtzOjQ6ImF0b20iO3M6NTA6Ii9tZWRyeGl2L2Vhcmx5LzIwMjIvMTAvMjIvMjAyMi4xMC4yMS4yMjI4MDM0OS5hdG9tIjt9czo4OiJmcmFnbWVudCI7czowOiIiO30=) 69. 69.Bal AM, Coombs GW, Holden MTG, Lindsay JA, Nimmo GR, Tattevin P, Skov RL. 2016. Genomic insights into the emergence and spread of international clones of healthcare-, community- and livestock-associated meticillin-resistant Staphylococcus aureus: Blurring of the traditional definitions. J Glob Antimicrob Resist 6:95–101. 70. 70.Hsu LY, Harris SR, Chlebowicz MA, Lindsay JA, Koh TH, Krishnan P, Tan TY, Hon PY, Grubb WB, Bentley SD, Parkhill J, Peacock SJ, Holden MTG. 2015. Evolutionary dynamics of methicillin-resistant Staphylococcus aureus within a healthcare system. Genome Biol 16:1–13. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1186/s13059-014-0572-2&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=25583448&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) 71. 71.Kong EF, Johnson JK, Jabra-Rizk MA. 2016. Community-Associated Methicillin-Resistant Staphylococcus aureus: An Enemy amidst Us. PLOS Pathog 12:e1005837. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1371/journal.ppat.1005837&link_type=DOI) 72. 72.Strauß L, Stegger M, Akpaka PE, Alabi A, Breurec S, Coombs G, Egyir B, Larsen AR, Laurent F, Monecke S, Peters G, Skov R, Strommenger B, Vandenesch F, Schaumburg F, Mellmann A. Origin, evolution, and global transmission of community-acquired Staphylococcus aureus ST8 [https://doi.org/10.1073/pnas.1702472114](https://doi.org/10.1073/pnas.1702472114). 73. 73.Wyllie DH, Walker AS, Miller R, Moore C, Williamson SR, Schlackow I, Finney JM, O’Connor L, Peto TEA, Crook DW. 2011. Decline of meticillin-resistant Staphylococcus aureus in Oxfordshire hospitals is strain-specific and preceded infection-control intensification. BMJ Open 1:e000160. [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NzoiYm1qb3BlbiI7czo1OiJyZXNpZCI7czoxMToiMS8xL2UwMDAxNjAiO3M6NDoiYXRvbSI7czo1MDoiL21lZHJ4aXYvZWFybHkvMjAyMi8xMC8yMi8yMDIyLjEwLjIxLjIyMjgwMzQ5LmF0b20iO31zOjg6ImZyYWdtZW50IjtzOjA6IiI7fQ==) 74. 74.Holden MTG, Hsu LY, Kurt K, Weinert LA, Mather AE, Harris SR, Strommenger B, Layer F, Witte W, De Lencastre H, Skov R, Westh H, Žemličková H, Coombs G, Kearns AM, Hill RLR, Edgeworth J, Gould I, Gant V, Cooke J, Edwards GF, McAdam PR, Templeton KE, McCann A, Zhou Z, Castillo-Ramírez S, Feil EJ, Hudson LO, Enright MC, Balloux F, Aanensen DM, Spratt BG, Fitzgerald JR, Parkhill J, Achtman M, Bentley SD, Nübel U. 2013. A genomic portrait of the emergence, evolution, and global spread of a methicillin-resistant Staphylococcus aureus pandemic. Genome Res 23:653–664. [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NjoiZ2Vub21lIjtzOjU6InJlc2lkIjtzOjg6IjIzLzQvNjUzIjtzOjQ6ImF0b20iO3M6NTA6Ii9tZWRyeGl2L2Vhcmx5LzIwMjIvMTAvMjIvMjAyMi4xMC4yMS4yMjI4MDM0OS5hdG9tIjt9czo4OiJmcmFnbWVudCI7czowOiIiO30=) 75. 75.Aires-de-Sousa M, Correia B, De Lencastre H, Alves V, Branca F, Cabral L, Clemente J, Daniel I, Faustino A, Ferreira E, Lameiras C, Lopes J, Marques J, Peres I, Ribeiro G, Sancho L, Santos O, Santos P, Vaz MT, Videira Z. 2008. Changing patterns in frequency of recovery of five methicillin-resistant Staphylococcus aureus clones in Portuguese hospitals: Surveillance over a 16-year period. J Clin Microbiol 46:2912–2917. [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MzoiamNtIjtzOjU6InJlc2lkIjtzOjk6IjQ2LzkvMjkxMiI7czo0OiJhdG9tIjtzOjUwOiIvbWVkcnhpdi9lYXJseS8yMDIyLzEwLzIyLzIwMjIuMTAuMjEuMjIyODAzNDkuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 76. 76.Gill JL, Hedge J, Wilson DJ, MacLean RC. 2021. Evolutionary Processes Driving the Rise and Fall of Staphylococcus aureus ST239, a Dominant Hybrid Pathogen. MBio [https://doi.org/10.1128/MBIO.02168-21](https://doi.org/10.1128/MBIO.02168-21). 77. 77.Wang M, Zheng Y, Mediavilla JR, Chen L, Kreiswirth BN, Song Y, Yang R, Du H. 2017. Hospital Dissemination of tst-1-Positive Clonal Complex 5 (CC5) Methicillin-Resistant Staphylococcus aureus. Front Cell Infect Microbiol 7. 78. 78.Goudarzi M, Razeghi M, Chirani AS, Fazeli M, Tayebi Z, Pouriran R. 2020. Characteristics of methicillin-resistant Staphylococcus aureus carrying the toxic shock syndrome toxin gene: high prevalence of clonal complex 22 strains and the emergence of new spa types t223 and t605 in Iran. New Microbes New Infect 36:100695. 79. 79.Brescó MS, Harris LG, Thompson K, Stanic B, Morgenstern M, O’Mahony L, Richards RG, Moriarty TF. 2017. Pathogenic mechanisms and host interactions in Staphylococcus epidermidis device-related infection. Front Microbiol 8:1401. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.3389/fmicb.2017.01401&link_type=DOI) 80. 80.Fey PD, Olson ME. 2010. Current concepts in biofilm formation of Staphylococcus epidermidis. Future Microbiol 5:917. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.2217/fmb.10.56&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=20521936&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000278582500013&link_type=ISI) 81. 81.Mack D. 1999. Molecular mechanisms of Staphylococcus epidermidis biofilm formation. J Hosp Infect 43:S113–S125. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1016/S0195-6701(99)90074-9&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=10658767&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) 82. 82.Harris LG, Murray S, Pascoe B, Bray J, Meric G, Magerios L, Wilkinson TS, Jeeves R, Rohde H, Schwarz S, De Lencastre H, Miragaia M, Rolo J, Bowden R, Jolley KA, Maiden MCJ, Mack D, Sheppard SK. 2016. Biofilm Morphotypes and Population Structure among Staphylococcus epidermidis from Commensal and Clinical Samples. PLoS One 11. 83. 83.Clauditz A, Resch A, Wieland KP, Peschel A, Götz F. 2006. Staphyloxanthin Plays a Role in the Fitness of Staphylococcus aureus and Its Ability To Cope with Oxidative Stress. Infect Immun 74:4950. [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MzoiaWFpIjtzOjU6InJlc2lkIjtzOjk6Ijc0LzgvNDk1MCI7czo0OiJhdG9tIjtzOjUwOiIvbWVkcnhpdi9lYXJseS8yMDIyLzEwLzIyLzIwMjIuMTAuMjEuMjIyODAzNDkuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 84. 84.Xue L, Chen YY, Yan Z, Lu W, Wan D, Zhu H. 2019. Staphyloxanthin: a potential target for antivirulence therapy. Infect Drug Resist 12:2151. 85. 85.Post V, Wahl P, Uçkay I, Ochsner P, Zimmerli W, Corvec S, Loiez C, Richards RG, Moriarty TF. 2014. Phenotypic and genotypic characterisation of Staphylococcus aureus causing musculoskeletal infections. Int J Med Microbiol 304:565–576. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1016/j.ijmm.2014.03.003&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=24768432&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) 86. 86.Rooijakkers SHM, Ruyken M, van Roon J, van Kessel KPM, van Strijp JAG, van Wamel WJB. 2006. Early expression of SCIN and CHIPS drives instant immune evasion by Staphylococcus aureus. Cell Microbiol 8:1282–1293. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1111/j.1462-5822.2006.00709.x&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=16882032&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000238884100007&link_type=ISI) 87. 87.De Haas CJC, Veldkamp KE, Peschel A, Weerkamp F, Van Wamel WJB, Heezius ECJM, Poppelier MJJG, Van Kessel KPM, Van Strijp JAG. 2004. Chemotaxis Inhibitory Protein of Staphylococcus aureus, a Bacterial Antiinflammatory Agent. J Exp Med 199:687. [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MzoiamVtIjtzOjU6InJlc2lkIjtzOjk6IjE5OS81LzY4NyI7czo0OiJhdG9tIjtzOjUwOiIvbWVkcnhpdi9lYXJseS8yMDIyLzEwLzIyLzIwMjIuMTAuMjEuMjIyODAzNDkuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 88. 88.Uçkay I, Teterycz D, Ferry T, Harbarth S, Lübbeke A, Emonet S, Chilcott M, Hoffmeyer P, Bernard L. 2009. Poor utility of MRSA screening to predict staphylococcal species in orthopaedic implant infections. J Hosp Infect 73:89–91. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1016/j.jhin.2009.06.016&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=19646786&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) 89. 89.Ferry T, Uçkay I, Vaudaux P, François P, Schrenzel J, Harbarth S, Laurent F, Bernard L, Vandenesch F, Etienne J, Hoffmeyer P, Lew D. 2010. Risk factors for treatment failure in orthopedic device-related methicillin-resistant Staphylococcus aureus infection. Eur J Clin Microbiol Infect Dis 29:171–180. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1007/s10096-009-0837-y&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=19946789&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000273745400006&link_type=ISI) 90. 90.Young BC, Wu C-H, Charlesworth J, Earle S, Price JR, Gordon NC, Cole K, Dunn L, Liu E, Oakley S, Godwin H, Fung R, Miller R, Knox K, Votintseva A, Quan TP, Tilley R, Scarborough M, Crook DW, Peto TE, Walker AS, Llewelyn MJ, Wilson DJ. 2021. Antimicrobial resistance determinants are associated with Staphylococcus aureus bacteraemia and adaptation to the healthcare environment: a bacterial genome-wide association study. Microb Genomics 7:700. 91. 91.Castillo-Ramírez S, Corander J, Marttinen P, Aldeljawi M, Hanage WP, Westh H, Boye K, Gulay Z, Bentley SD, Parkhill J, Holden MT, Feil EJ. 2012. Phylogeographic variation in recombination rates within a global clone of methicillin-resistant Staphylococcus aureus. Genome Biol 13:1–13. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1186/gb-2012-13-1-r1&link_type=DOI) 92. 92.Earle SG, Wu C-H, Charlesworth J, Stoesser N, Gordon NC, Walker TM, Spencer CCA, Iqbal Z, Clifton DA, Hopkins KL, Woodford N, Smith EG, Ismail N, Llewelyn MJ, Peto TE, Crook DW, McVean G, Walker AS, Wilson DJ. 2016. Identifying lineage effects when controlling for population structure improves power in bacterial association studies. Nat Microbiol 1:16041. 93. 93.Collins C, Didelot X. 2018. A phylogenetic method to perform genome-wide association studies in microbes that accounts for population structure and recombination. PLoS Comput Biol 14. 94. 94.Taylor AJ, Méric G, Yahara K, Pascoe B, Mageiros L, Mourkas E, Calland JK, Puranen S, Hitchings MD, Jolley KA, Kobras CM, Williams NJ, Vliet AHM van, Parkhill J, Maiden MCJ, Corander J, Hurst LD, Falush D, Didelot X, Kelly DJ, Sheppard SK. 2022. A two-hit epistasis model prevents core genome disharmony in recombining bacteria. bioRxiv 2021.03.15.435406. 95. 95.Yokoyama M, Stevens E, Laabei M, Bacon L, Heesom K, Bayliss S, Ooi N, O’Neill AJ, Murray E, Williams P, Lubben A, Reeksting S, Meric G, Pascoe B, Sheppard SK, Recker M, Hurst LD, Massey RC. 2018. Epistasis analysis uncovers hidden antibiotic resistance-associated fitness costs hampering the evolution of MRSA. Genome Biol 19. 96. 96.Paterson GK, Harrison EM, Murray GGR, Welch JJ, Warland JH, Holden MTG, Morgan FJE, Ba X, Koop G, Harris SR, Maskell DJ, Peacock SJ, Herrtage ME, Parkhill J, Holmes MA. 2015. Capturing the cloud of diversity reveals complexity and heterogeneity of MRSA carriage, infection and transmission. Nat Commun 2015 61 6:1–10. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1038/ncomms6945&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=25649611&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) 97. 97.Didelot X, Walker AS, Peto TE, Crook DW, Wilson DJ. 2016. Within-host evolution of bacterial pathogens. Nat Rev Microbiol. Nature Publishing Group. 98. 98.Golubchik T, Batty EM, Miller RR, Farr H, Young BC, Larner-Svensson H, Fung R, Godwin H, Knox K, Votintseva A, Everitt RG, Street T, Cule M, Ip CLC, Didelot X, Peto TEA, Harding RM, Wilson DJ, Crook DW, Bowden R. 2013. Within-Host Evolution of Staphylococcus aureus during Asymptomatic Carriage. PLoS One 8. 99. 99.Young BC, Golubchik T, Batty EM, Fung R, Larner-Svensson H, Votintseva AA, Miller RR, Godwin H, Knox K, Everitt RG, Iqbal Z, Rimmer AJ, Cule M, Ip CLC, Didelot X, Harding RM, Donnelly P, Peto TE, Crook DW, Bowden R, Wilsona DJ. 2012. Evolutionary dynamics of Staphylococcus aureus during progression from carriage to disease. Proc Natl Acad Sci 109:4550–4555. [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NDoicG5hcyI7czo1OiJyZXNpZCI7czoxMToiMTA5LzEyLzQ1NTAiO3M6NDoiYXRvbSI7czo1MDoiL21lZHJ4aXYvZWFybHkvMjAyMi8xMC8yMi8yMDIyLjEwLjIxLjIyMjgwMzQ5LmF0b20iO31zOjg6ImZyYWdtZW50IjtzOjA6IiI7fQ==) 100.100.Kahl BC. 2014. Small colony variants (SCVs) of Staphylococcus aureus –A bacterial survival strategy. Infect Genet Evol 21:515–522. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1016/j.meegid.2013.05.016.&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=23722021&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) 101.101.Loss G, Simões PM, Valour F, Cortês MF, Gonzaga L, Bergot M, Trouillet-Assant S, Josse J, Diot A, Ricci E, Vasconcelos AT, Laurent F. 2019. Staphylococcus aureus Small Colony Variants (SCVs): News From a Chronic Prosthetic Joint Infection. Front Cell Infect Microbiol 9:363. 102.102.Peyrusson F, Varet H, Nguyen TK, Legendre R, Sismeiro O, Coppée JY, Wolz C, Tenson T, Van Bambeke F. 2020. Intracellular Staphylococcus aureus persisters upon antibiotic exposure. Nat Commun 2020 111 11:1–14. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1038/s41467-019-13889-6&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=31911652&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) 103.103.Lechner S, Lewis K, Bertram R. 2012. Staphylococcus aureus Persisters Tolerant to Bactericidal Antibiotics. Microb Physiol 22:235–244. 104.104.Pascoe B, Dams L, Wilkinson TS, Harris LG, Bodger O, Mack D, Davies AP. 2014. Dormant cells of Staphylococcus aureus are resuscitated by spent culture supernatant. PLoS One 9. 105.105.Conlon BP, Rowe SE, Gandt AB, Nuxoll AS, Donegan NP, Zalis EA, Clair G, Adkins JN, Cheung AL, Lewis K. 2016. Persister formation in Staphylococcus aureus is associated with ATP depletion. Nat Microbiol 1. 106.106.Bankevich A, Nurk S, Antipov D, Gurevich AA, Dvorkin M, Kulikov AS, Lesin VM, Nikolenko SI, Pham S, Prjibelski AD, Pyshkin A V., Sirotkin A V., Vyahhi N, Tesler G, Alekseyev MA, Pevzner PA. 2012. SPAdes: A New Genome Assembly Algorithm and Its Applications to Single-Cell Sequencing. J Comput Biol 19:455–477. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1089/cmb.2012.0021&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=22506599&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) 107.107.Katoh K, Standley DM. 2013. MAFFT Multiple Sequence Alignment Software Version 7: Improvements in Performance and Usability. Mol Biol Evol 30:772–780. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1093/molbev/mst010&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=23329690&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000317002300004&link_type=ISI) 108.108.Nguyen L-T, Schmidt HA, von Haeseler A, Minh BQ. 2015. IQ-TREE: A Fast and Effective Stochastic Algorithm for Estimating Maximum-Likelihood Phylogenies. Mol Biol Evol 32:268–274. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1093/molbev/msu300&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=25371430&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) 109.109.Hoang DT, Chernomor O, von Haeseler A, Minh BQ, Vinh LS. 2018. UFBoot2: Improving the Ultrafast Bootstrap Approximation. Mol Biol Evol 35:518–522. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1093/molbev/msx281&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=29077904&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) 110.110.Argimón S, Abudahab K, Goater RJE, Fedosejev A, Bhai J, Glasner C, Feil EJ, Holden MTG, Yeats CA, Grundmann H, Spratt BG, Aanensen DM. 2016. Microreact: visualizing and sharing data for genomic epidemiology and phylogeography. Microb Genomics 2. 111.111.Jolley KA, Bray JE, Maiden MCJ. 2018. Open-access bacterial population genomics: BIGSdb software, the [PubMLST.org](http://PubMLST.org) website and their applications. Wellcome open Res 3. 112.112.Petit RA, Read TD. 2018. Staphylococcus aureus viewed from the perspective of 40,000+ genomes. PeerJ 2018:e5261. 113.113.Bartels MD, Petersen A, Worning P, Nielsen JB, Larner-Svensson H, Johansen HK, Andersen LP, Jarløv JO, Boye K, Larsen AR, Westh H. 2014. Comparing Whole-Genome Sequencing with Sanger Sequencing for spa Typing of Methicillin-Resistant Staphylococcus aureus. J Clin Microbiol 52:4305. [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MzoiamNtIjtzOjU6InJlc2lkIjtzOjEwOiI1Mi8xMi80MzA1IjtzOjQ6ImF0b20iO3M6NTA6Ii9tZWRyeGl2L2Vhcmx5LzIwMjIvMTAvMjIvMjAyMi4xMC4yMS4yMjI4MDM0OS5hdG9tIjt9czo4OiJmcmFnbWVudCI7czowOiIiO30=) 114.114.Chen L, Zheng D, Liu B, Yang J, Jin Q. 2016. VFDB 2016: Hierarchical and refined dataset for big data analysis - 10 years on. Nucleic Acids Res 44:D694–D697. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1093/nar/gkv1239&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=26578559&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) 115.115.Feldgarden M, Brover V, Haft DH, Prasad AB, Slotta DJ, Tolstoy I, Tyson GH, Zhao S, Hsu C-H, McDermott PF, Tadesse DA, Morales C, Simmons M, Tillman G, Wasilenko J, Folster JP, Klimke W. 2019. Using the NCBI AMRFinder Tool to Determine Antimicrobial Resistance Genotype-Phenotype Correlations Within a Collection of NARMS Isolates. bioRxiv 550707. 116.116.Alcock BP, Raphenya AR, Lau TTY, Tsang KK, Bouchard M, Edalatmand A, Huynh W, Nguyen A-L V, Cheng AA, Liu S, Min SY, Miroshnichenko A, Tran H-K, Werfalli RE, Nasir JA, Oloni M, Speicher DJ, Florescu A, Singh B, Faltyn M, Hernandez-Koutoucheva A, Sharma AN, Bordeleau E, Pawlowski AC, Zubyk HL, Dooley D, Griffiths E, Maguire F, Winsor GL, Beiko RG, Brinkman FSL, Hsiao WWL, Domselaar G V, McArthur AG. 2019. CARD 2020: antibiotic resistome surveillance with the comprehensive antibiotic resistance database. Nucleic Acids Res [https://doi.org/10.1093/nar/gkz935](https://doi.org/10.1093/nar/gkz935). 117.117.Altschul SF, Gish W, Miller W, Myers EW, Lipman DJ. 1990. Basic local alignment search tool. J Mol Biol 215:403–410. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1006/jmbi.1990.9999&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=2231712&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=A1990ED16700008&link_type=ISI) 118.118.EFSA/ECDC. 2019. European Food Safety Authority and European Centre for Disease Prevention and Control. The European Union summary report on antimicrobial resistance in zoonotic and indicator bacteria from humans, animals and food in 2017. EFSA J 17:5598. 119.119.Herbert S, Ziebandt AK, Ohlsen K, Schäfer T, Hecker M, Albrecht D, Novick R, Götz F. 2010. Repair of Global Regulators in Staphylococcus aureus 8325 and Comparative Analysis with Other Clinical Isolates. Infect Immun 78:2877. [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MzoiaWFpIjtzOjU6InJlc2lkIjtzOjk6Ijc4LzYvMjg3NyI7czo0OiJhdG9tIjtzOjUwOiIvbWVkcnhpdi9lYXJseS8yMDIyLzEwLzIyLzIwMjIuMTAuMjEuMjIyODAzNDkuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 120.120.Stepanović S, Vuković D, Hola V, Di Bonaventura G, Djukić S, Ćirković I, Ruzicka F. 2007. Quantification of biofilm in microtiter plates: overview of testing conditions and practical recommendations for assessment of biofilm production by staphylococci. APMIS 115:891–899. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1111/j.1600-0463.2007.apm_630.x&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=17696944&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000248668000001&link_type=ISI) 121.121.Ignasiak K, Maxwell A. 2017. Galleria mellonella (greater wax moth) larvae as a model for antibiotic susceptibility testing and acute toxicity trials. BMC Res Notes 10:428. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1186/s13104-017-2757-8&link_type=DOI) 122.122.Desbois AP, Coote PJ. 2011. Wax moth larva (Galleria mellonella): an in vivo model for assessing the efficacy of antistaphylococcal agents. J Antimicrob Chemother 66:1785–1790. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1093/jac/dkr198&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=21622972&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000292696400017&link_type=ISI) 123.123.Croucher NJ, Page AJ, Connor TR, Delaney AJ, Keane JA, Bentley SD, Parkhill J, Harris SR. 2015. Rapid phylogenetic analysis of large samples of recombinant bacterial whole genome sequences using Gubbins. Nucleic Acids Res 43:e15–e15. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1093/nar/gku1196&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=25414349&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) 124.124.Mageiros L, Méric G, Bayliss SC, Pensar J, Pascoe B, Mourkas E, Calland JK, Yahara K, Murray S, Wilkinson TS, Williams LK, Hitchings MD, Porter J, Kemmett K, Feil EJ, Jolley KA, Williams NJ, Corander J, Sheppard SK. 2021. Genome evolution and the emergence of pathogenicity in avian Escherichia coli. Nat Commun 12. 125.125.Mourkas E, Florez-Cuadrado D, Pascoe B, Calland JK, Bayliss SC, Mageiros L, Méric G, Hitchings MD, Quesada A, Porrero C, Ugarte-Ruiz M, Gutiérrez-Fernández J, Domínguez L, Sheppard SK. 2019. Gene pool transmission of multidrug resistance among Campylobacter from livestock, sewage and human disease. Environ Microbiol [https://doi.org/10.1111/1462-2920.14760](https://doi.org/10.1111/1462-2920.14760). 126.126.Schliep KP. 2011. phangorn: phylogenetic analysis in R. Bioinformatics 27:592–593. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1093/bioinformatics/btq706&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=21169378&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000287246000025&link_type=ISI) 127.127.Connor TR, Loman NJ, Thompson S, Smith A, Southgate J, Poplawski R, Bull MJ, Richardson E, Ismail M, Thompson SE-, Kitchen C, Guest M, Bakke M, Sheppard SK, Pallen MJ. 2016. CLIMB (the Cloud Infrastructure for Microbial Bioinformatics): an online resource for the medical microbiology community. Microb Genomics 2. 128.128.Grüning B, Dale R, Sjödin A, Chapman BA, Rowe J, Tomkins-Tinch CH, Valieris R, Köster J. 2018. Bioconda: sustainable and comprehensive software distribution for the life sciences. Nat Methods 15:475–476. 129.129.Jolley KA, Maiden MCJ. 2010. BIGSdb: Scalable analysis of bacterial genome variation at the population level. BMC Bioinformatics 11:595. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1186/1471-2105-11-595&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=21143983&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2022%2F10%2F22%2F2022.10.21.22280349.atom)