Skip to main content
Antimicrobial Agents and Chemotherapy logoLink to Antimicrobial Agents and Chemotherapy
. 2014 Apr;58(4):1907–1917. doi: 10.1128/AAC.02653-13

Extensive Genetic Diversity Identified among Sporadic Methicillin-Resistant Staphylococcus aureus Isolates Recovered in Irish Hospitals between 2000 and 2012

Peter M Kinnevey a, Anna C Shore a,b, Grainne I Brennan a,c, Derek J Sullivan a, Ralf Ehricht d, Stefan Monecke d,e, David C Coleman a,
PMCID: PMC4023797  PMID: 24395241

Abstract

Clonal replacement of predominant nosocomial methicillin-resistant Staphylococcus aureus (MRSA) strains has occurred several times in Ireland during the last 4 decades. However, little is known about sporadically occurring MRSA in Irish hospitals or in other countries. Eighty-eight representative pvl-negative sporadic MRSA isolates recovered in Irish hospitals between 2000 and 2012 were investigated. These yielded unusual pulsed-field gel electrophoresis and antibiogram-resistogram typing patterns distinct from those of the predominant nosocomial MRSA clone, ST22-MRSA-IV, during the study period. Isolates were characterized by spa typing and DNA microarray profiling for multilocus sequence type (MLST) clonal complex (CC) and/or sequence type (ST) and SCCmec type assignment, as well as for detection of virulence and antimicrobial resistance genes. Conventional PCR-based SCCmec subtyping was undertaken when necessary. Extensive diversity was detected, including 38 spa types, 13 MLST-CCs (including 18 STs among 62 isolates assigned to STs), and 25 SCCmec types (including 2 possible novel SCCmec elements and 7 possible novel SCCmec subtypes). Fifty-four MLST-spa-SCCmec type combinations were identified. Overall, 68.5% of isolates were assigned to nosocomial lineages, with ST8-t190-MRSA-IID/IIE ± SCCM1 predominating (17.4%), followed by CC779/ST779-t878-MRSA-ψSCCmec-SCC-SCCCRISPR (7.6%) and CC22/ST22-t032-MRSA-IVh (5.4%). Community-associated clones, including CC1-t127/t386/t2279-MRSA-IV, CC59-t216-MRSA-V, CC8-t008-MRSA-IVa, and CC5-t002/t242-MRSA-IV/V, and putative animal-associated clones, including CC130-t12399-MRSA-XI, ST8-t064-MRSA-IVa, ST398-t011-MRSA-IVa, and CC6-t701-MRSA-V, were also identified. In total, 53.3% and 47.8% of isolates harbored genes for resistance to two or more classes of antimicrobial agents and two or more mobile genetic element-encoded virulence-associated factors, respectively. Effective ongoing surveillance of sporadic nosocomial MRSA is warranted for early detection of emerging clones and reservoirs of virulence, resistance, and SCCmec genes.

INTRODUCTION

Staphylococcus aureus colonizes the anterior nares of approximately 30% of the human population; it can give rise to a wide range of infections of skin and soft tissues, bones, joints, and prosthetic implants and can be responsible for a variety of toxinoses caused by specific toxins such as toxic shock toxin, enterotoxins, exfoliative toxins, and Panton-Valentine leukocidin (1). Staphylococcus aureus can evolve to methicillin-resistant S. aureus (MRSA) upon acquisition of a large staphylococcal chromosomal cassette (SCC) element harboring either the methicillin resistance gene mecA or mecC (SCCmec), both of which encode a modified penicillin-binding protein, PBP 2a (24).

Within SCCmec, mecA or mecC forms part of the mec gene complex, which may also harbor the mec regulatory genes mecI and mecR1, as well as insertion sequences and, in some instances, blaZ (13). Based on various combinations and truncations of the mec complex genes, five classes of the mec gene complex (A, B, C1, C2, and E) have been identified in MRSA (3, 5). In addition, each SCCmec element also harbors a chromosome cassette recombinase (ccr) gene complex, consisting of ccrA and ccrB together or ccrC; these genes encode polypeptides that catalyze site- and orientation-specific integration and excision of SCCmec into orfX within the S. aureus chromosome (1, 6). Seven types of ccr gene complex (types 1 to 5, 7, and 8) have been described to occur in MRSA, each with a different combination of ccrA and ccrB or ccrC alleles (5, 7). SCC elements that carry ccr genes but lack mec genes have also been described, as well as pseudo (ψ) SCCmec and SCC elements that lack ccr genes, individual SCCmec elements with multiple ccr genes, and composite islands (CIs) consisting of two or more elements (5).

Eleven SCCmec types (I to XI) have been described to date for MRSA, each with a different combination of mec class and ccr type (3). Numerous SCCmec subtypes have also been described for MRSA which differ from SCCmec types based on DNA sequence variation or the presence or absence of mobile genetic elements (MGEs) in the joining (J) regions, which are located outside the ccr and mec complexes (7). MRSA isolates often exhibit resistance to a range of antimicrobial agents that can be due to the carriage of multiple antimicrobial resistance genes located on MGEs, including transposons, plasmids, and SCC or SCCmec elements (8, 9).

The first report of MRSA appeared in the literature in 1961, shortly after the introduction of methicillin into clinical use; just 10 years later, in 1971, MRSA was first reported in Irish hospitals (10, 11). Following a major increase in the prevalence of MRSA in Irish hospitals in the late 1970s and during the 1980s and 1990s, it has now been endemic for more than 3 decades (1217). Since 1999, the prevalence rate of MRSA among S. aureus strains causing bloodstream infections (BSIs) in Ireland has been monitored by the European Antimicrobial Resistance Surveillance Network (EARS-Net). Annual rates of MRSA among S. aureus from BSIs in Ireland reached 42% (592 MRSA isolates among 1,412 S. aureus isolates) in 2006, the highest level reported to date, and declined in recent years, with a rate of 22.8% (242 MRSA isolates among 1,060 S. aureus isolates) reported for 2012 (18, 19).

Clonal replacement of predominant nosocomial MRSA strains has occurred several times in Ireland during the last 4 decades (20). The different MRSA lineages that predominated in Irish hospitals at different time periods have been well characterized, including ST250-MRSA-I/I-pls (where ST is multilocus sequence type) in the 1970s and early 1980s, ST239-MRSA-III/III-pI258/Tn554 in the middle to late 1980s and early 1990s, and ST8-MRSA-IIA-IIE throughout the 1990s, together with ST36-MRSA-II and ST22-MRSA-IV in the late 1990s; since 2002, the ST22-MRSA-IV clone has dominated (17, 20). Prior to 1999, ST22-MRSA-IV was detected only sporadically among MRSA isolates in Ireland, but by 2003, it accounted for 80% of MRSA BSIs, and despite a decline in the proportion of S. aureus infections due to MRSA in recent years, it has continued to account for 70 to 80% of MRSA BSIs each year (19).

While a limited number of sporadically occurring MRSA clones from patients in Irish hospitals in the 1980s and 1990s have been characterized using multilocus sequence typing (MLST), SCCmec typing, and DNA microarray analysis, e.g., ST5-MRSA-II and ST247-MRSA-Ia, there have been no systematic detailed studies of the genetic diversity of sporadic MRSA strains in Ireland (20, 21). These account for approximately 20 to 30% of MRSA BSIs in Ireland each year, as well as being identified each year among non-BSI isolates submitted to the Irish National MRSA Reference Laboratory (NMRSARL) from patients in hospitals with a variety of infections and from patient and environmental screening samples (19). In fact, the number of sporadic MRSA isolates identified among BSIs in Ireland increased from 12.1% in 2005 to 23.1% in 2011 (19). Numerous studies have shown that many MRSA clones that occur sporadically or not at all in one geographic region are often prevalent in another region and vice versa (20, 22, 23). However, previous studies that have investigated sporadic MRSA populations are limited in terms of sample size and/or depth of analysis (2427).

Due to the potential of sporadic MRSA strains to replace currently dominant MRSA clones and because they account for a significant proportion of MRSA infections in Ireland each year, it is essential that populations of new and emerging MRSA strains be monitored. In addition, sporadic MRSA strains may constitute a significant potential reservoir for virulence and resistance genes located on MGEs, in particular SCCmec elements. Therefore, the present study investigated the genotypes, SCCmec types, and virulence and resistance genes within 88 MRSA isolates representative of 1,663 pvl-negative sporadically occurring MRSA isolates from patients in Irish hospitals between 2000 and 2012. Isolates were investigated using spa typing, MLST, SCCmec typing, and DNA microarray profiling. The 88 sporadic MRSA isolates were selected at the NMRSARL based on unusual antibiogram-resistogram (AR) and/or pulsed-field gel electrophoresis (PFGE) typing patterns which were different from that of the endemic strain that predominated in Irish hospitals during the study period, i.e., ST22-MRSA-IV. All pvl-positive MRSA isolates from Irish hospitals and community sources submitted to the NMRSARL for examination during the same period have been investigated as part of a separate study (28).

MATERIALS AND METHODS

Bacterial isolates.

MRSA isolates identified by the NMRSARL were deemed to be sporadic if they exhibited unusual AR and/or PFGE typing patterns which were different from those of the endemic strain in Irish hospitals during the study period, i.e., ST22-MRSA-IV. Unusual AR type patterns included those that were different from previously described ST22-MRSA-IV AR (AR06) type patterns. Unusual PFGE patterns were identified using the criteria of Tenover et al. (29) and differed from PFGE patterns previously identified among ST22-MRSA-IV isolates by ≥7 PFGE bands. Using these criteria, a total of 1,663 pvl-negative sporadic MRSA isolates were identified by the NMRSARL from patients in Irish hospitals between 2000 and 2012. Of the 1,663 isolates, 841 were investigated and determined to be pvl negative either by PCR, as described previously (30), or using an in-house real-time PCR assay. The remaining 822 isolates were not investigated for pvl, as they yielded AR and/or PFGE typing patterns indicative of strains not previously associated with pvl, e.g., AR13 and AR14, and therefore a pvl-negative status was inferred for these (30, 31). All pvl-positive isolates recovered in Ireland during the study period between 2000 and 2012 were investigated as part of a separate study (28). Eighty-seven of the 1,663 pvl-negative sporadic MRSA isolates, representative of approximately 5% of sporadic pvl-negative MRSA isolates identified each year from patients in Irish hospitals during the 12-year study period (see Table S1 in the supplemental material) and representing as diverse a range as possible of AR and/or PFGE typing patterns, were selected for detailed investigation. In addition, one pvl-negative MRSA isolate recovered from a patient in the community but who had previously been hospitalized on several occasions was also included for investigation. This isolate harbored mecC (SCCmec type XI) and was included because this clone was recovered sporadically in two patients in two separate hospitals in Ireland in 2010 (2) and has recently been reported in several other European countries (32). The majority of the MRSA isolates (73.8% [65/88]) selected for study were recovered from infections (89% [58/65] BSIs and 10.8% [7/65] skin and soft tissue infections [SSTIs]), 11% (10/88) were colonizing isolates from patient screening, and no information was available for the remainder (14.8% [13/88]). Isolates were identified as S. aureus, and methicillin resistance was confirmed as described previously (30). Isolates were stored on Protect beads (Technical Service Consultants Limited, Heywood, United Kingdom) at −70°C prior to subsequent detailed analysis.

DNA microarray analysis.

The 88 sporadic MRSA isolates were investigated by DNA microarray profiling using the StaphyType kit (Alere Technologies GmbH, Jena, Germany). The StaphyType kit consists of individual DNA microarrays mounted in 8-well microtiter strips which detect 333 S. aureus gene sequences and alleles, including species-specific, antimicrobial resistance and virulence-associated genes, SCCmec genes, typing markers, and a staining control (33, 34). ArrayMate software (version 2012-01-18) (Alere Technologies) was used to analyze data generated by the microarray system and to assign isolates to inferred sequence types (STs) and/or clonal complexes (CCs) by comparing the DNA microarray profile results of test isolates to microarray profiles of an extensive range of reference strains held in the ArrayMate database that have been previously typed by MLST (33, 34). The DNA array can assign all isolates investigated to the correct MLST clonal complex (CC) with a 98% correlation with STs assigned by MLST (21). Genomic DNA for use with the DNA microarray was extracted from all isolates by enzymatic lysis using the buffers and solutions provided with the StaphyType kit and the Qiagen DNeasy blood and tissue kit (Qiagen, Crawley, West Sussex, United Kingdom). The primers, probes, and protocols for this DNA microarray system have been described in detail previously (34).

In order to visualize the similarities between the 88 sporadic isolates investigated (although not necessarily true phylogenetic relationships), a network tree was constructed using the complete DNA microarray hybridization profile data of the isolates using the software program SplitsTree, version 4.11.3 (35), as described previously (1). Array hybridization profiles of the isolates were converted into a series of strings of letters that can be handled by the software as sequences. For comparison, array profiles of 3,139 MRSA isolates representative of MRSA isolates globally that were characterized in a previous study were included for comparison (1).

Molecular typing.

Genomic DNA for spa typing, MLST, and SCCmec typing was extracted from each isolate using enzymatic lysis and the DNeasy blood and tissue kit (Qiagen) according to the manufacturer's instructions. Unless otherwise stated, PCRs were performed using GoTaq Flexi DNA polymerase (Promega Corporation, Madison, WI) according to the manufacturer's instructions and using the published methods for each PCR protocol as described below. PCR amplifications were performed in a G-storm GS1 thermocycler (Applied Biosystems, Foster City, CA). PCR products were visualized by conventional agarose gel electrophoresis and purified with the GenElute PCR cleanup kit (Sigma-Aldrich Ireland Ltd., Arklow, County Wicklow, Ireland).

All isolates underwent spa typing using the primers and thermal cycling conditions described by the European Network of Laboratories for Sequence Based Typing of Microbial Pathogens (SeqNet [http://www.seqnet.org]). Sequencing was performed commercially by Geneservice Limited (Source Bioscience, Dublin, Ireland) using an ABI 3730xl Sanger sequencing platform. Sequences were analyzed and were assigned to spa types using the Ridom StaphType software program, version 1.3 (Ridom GmbH, Wuerzburg, Germany) (36).

Although all isolates were assigned to STs and/or CCs using the DNA microarray, this system became available only during the latter half of the study. Prior to 2006, MLST had been performed on a subset of sporadic MRSA isolates (n = 27) representative of different spa types (Table 1). MLST was performed as described previously (37), sequences were analyzed using BioNumerics software, version 7.1 (Applied Maths, Ghent, Belgium), and alleles and STs were assigned using the MLST database (http://www.mlst.net/).

TABLE 1.

Molecular characteristics of 88 sporadic MRSA isolates recovered from patients in Irish hospitals between 2000 and 2012

Isolate reference number(s)a CC/ST-spa type SCCmec type/description (n) agr type Capsule type IEC type (n)b Antimicrobial resistance genes (n) Virulence genes (n) Locations where similar isolates have been reported (reference)
52, 88 CC/ST1-t2279 IVa (2) III 8 D (1), E (1) blaZ (2), fusC (1), sdrM (2) sea (1), sek (2), seq (2), seh (2) Western Australia (1)
62, 86 CC1-t2279c IVa and SCCfus (2) III 8 D blaZ (2), fusC (2), sdrM (2) sea (2), sek (2), seq (2), seh (2) Malta (68)
48 CC1-t386 IVa (1) III 8 E blaZ, erm(C), aphA3, sat, sdrM seh Germany (1)
7 CC1-t127 IVa (1) III 8 D blaZ, sdrM sea (1), sek, seq, seh Western Australia (1)
38, 41 CC1-t127 IVa and SCCfus (2) III 8 D blaZ (2), erm(A) (2), fusC (2), sdrM (2) sea (2), sek (2), seq (2), seh (2) Malta (68)
29 CC1/ST1336-t127c IVc (1) III 8 D blaZ, tet(K), sdrM sea, seb, sek, seq, seh None
66 CC1/ST1115-t127c IVa (1) III 8 E blaZ, erm(C), aphA3, sat, tet(K), sdrM seh None
8, 19 CC/ST5-t045 II (2) II 5 D (1), nege (1) blaZ (2), erm(A) (2), aadD (2), sdrM (2), fosB (2), qac(A) (1) tst (2), sea (1), egc (2), sed (1), sej (1), ser (1) None
32, 68 CC/ST5-t242c VT (2) II 5 B blaZ (2), aacA-aphD (2), sdrM (2), fosB (2) egc (2) USA (69)
85 CC5-t002 VT (harboring ccrC2) (1) II 5 G blaZ, aacA-aphD, sdrM, fosB egc, sep, sed, sej, ser None
75 CC5-t002 II (1) II 5 F erm(A), aacA-aphD, aadD, mupA, sdrM, fosB, qac(C) egc, sep Pandemic (1)
72 CC5-t002 IVc (1) II 5 F blaZ, sdrM, fosB egc, sep, sed, sej, ser Denmark (38)
79 CC5-t002 IV (nonsubtypeable) (1) II 5 F blaZ, sdrM, fosB egc, sep Pandemic (1)
64 CC5/ST930c,d-t002b IV (nonsubtypeable) (1) II 5 B blaZ, erm(C), aacA-aphD, sdrM, fosB egc None
21 CC5/ST100-t002c,d Novel 1 (mecA only detected) (1) II 5 E blaZ, aacA-aphD, sdrM, fosB egc None
71 CC5-t067 IV (nonsubtypeable) & ccrB4 (1) II 5 F blaZ, msrA, mph(C), aacA-aphD, aadD, aphA3, sat, mupA, sdrM, fosB, qac(C) egc, sep Spain (70)
55 CC5-t088 VT (1) II 5 D erm(C), sdrM, fosB sea, egc, sed, sej, ser None
14 CC/ST5-t109c,d I and ccrC (1) II 5 B blaZ, erm(A), aacA-aphD, aphA3, sat, tet(K), sdr(M), fosB, qac(A) egc None
67 CC5/ST1435d-t242c VT (1) II 5 B blaZ, aacA-aphD, sdrM,fosB egc None
78 CC5-t442 V (harboring ccrC8) (1) II 5 E blaZ, aacA-aphD, sdrM, fosB egc, sed, sej, ser Australia (71)
37 CC/ST5-t463 II (1) II 5 A blaZ, erm(A), aadD, sdrM, fosB tst, sea, egc, sed, sej, ser None
40 CC5-t1781 IVa (1) II 5 G blaZ, msrA, mph(C), aphA3, sat, sdrM, fosB, qac(C) egc, sep, sej, ser Germany, Canada (spa.ridom.de)
82 CC6-t701 IVh (1) I 8 E blaZ, sdrM, fosB neg Australia, Abu Dhabi, Hong Kong (72)
2, 6, 9, 13, 17, 24, 27, 50, 15, 16 CC/ST8-t190 IID and SCCM1 (10) I 5 D (8), neg (2) blaZ (10), erm(A) (10), aacA-aphD (10), aadD (1), aphA3 (4), sat (4), fusB (1), tet(K) (1), sdrM (10), cat (1), fosB (10), qacA (9) sea (8), neg (2) None
5, 12, 60, 43 CC/ST8-t190 IIE and SCCM1 (4) I 5 D (3), neg (1) blaZ (4), erm(A) (4), aacA-aphD (4), aadD (1), aphA3 (3), sat (3), mupA (1), sdrM (4), fosB (4), qacA (4) sea (3), neg (1) None
26 CC/ST8-t190c IID (1) I 5 A erm(A), aacA-aphD, aphA3, sat, sdrM, fosB tst, sea None
56 CC/ST8-t190c VI (1) I 5 E blaZ, erm(A), aphA3, sat, sdrM, fosB neg (1) None
39, 53 CC8-t008 IVa (2) I 5 B (1), neg (1) blaZ (2), erm(A) (1), msrA (1), mph(C) (1), aphA3 (1), sat (1), sdrM (2), fosB (2) ACME-arc (2) USA (1)
42 CC/ST8-t064c IVa (1) Ì 5 E blaZ, erm(C), sdrM, fosB seb, sek, seq USA, Switzerland (73, 74)
35 CC/ST8-t4268 IVd (1) I 5 D blaZ, erm(C), aacA-aphD, dfrS1, tet(M), sdrM, fosB, qac(A), qac(C) sea, neg None
61 CC/ST8-t1209c IIIB (1) I 8 neg blaZ, erm(A), aacA-aphD, aadD, tet(M), sdrM, fosB, qac(A) sek, seq None
31 CC8/ST239-t030c IIIB (1) I 8 neg blaZ, erm(A), tet(M), sdrM, fosB, qac(A) sek, seq Pandemic (75)
22 CC8/ST239-t037 IIIA (1) I 8 D blaZ, erm(A), aacA-aphD, aphA3, sat, tet(M), sdrM, fosB, qac(A) sea, sek, seq Pandemic (1)
18 CC8/ST239-t037c III and SCChg (1) I 8 D blaZ, erm(A), aacA-aphD, tet(K), tet(M), sdrM, aphA3, sat, fosB, qac(A) sea, sek, seq Pandemic (1)
11, 20, 25, 10, 28 CC/ST22-t032c IVh (5) I 5 B (3), neg (2) blaZ (5), erm(C) (4), aphA3 (1), sat (1), qac(A) (1) egc (4), sec (2), sel (2) Pandemic (1)
33 CC/ST22-t032c IVg (1) I 5 B blaZ, erm(C) egc, sec, sel Pandemic (1)
30 CC/ST22-t022c IVh (1) I 5 B blaZ, erm(C) egc, sec, sel Pandemic (1)
36 CC/ST22-t2951 IVh (1) I 5 B blaZ, erm(C), lnu (A), aacA-aphD, aadD, mupA, cat, fosB, qac(C) egc None
51 CC22-t1802 IVh (1) I 5 B blaZ, erm(C), fosB egc, sed None
57, 1, 3, 4 CC/ST36-t018c II (4) III 8 A (4) blaZ (4), erm(A) (4), aadD (2), sdrM (4), fosB (4) tst (3), sea (4), sed (1), egc (4) UK (76)
63 CC/ST36-t018c II and ccrC (1) III 5 A blaZ, erm(A), aacA-aphD, aadD, sdrM, fosB tst, sea, egc None
34 CC30/ST36-t012 II (1) III 5 B blaZ, erm(A), aacA-aphD, aadD, mupA, sdrM, fosB tst, egc UK (76)
23, 45, 54, 87 CC/ST45-t727 IVa (4) IV 8 B (2), neg (2) blaZ (4), erm(C) (2), fusB (4), mupA (1), tet(M) (2), sdrM (4), fosB (1) egc (4) Hong Kong, Australia (1)
58 CC/ST45-t727c Novel 2 (mecA only detected) (1) IV 8 neg blaZ, sdrM egc None
77 CC/ST45-t132 IVa (1) I 8 B blaZ, sdrM egc Germany, Belgium (1)
73 CC/ST45-t026 IVa (1) I 8 B blaZ, erm(C), sdrM egc, sec, sel Germany, Belgium (1)
81 CC/ST45-t065 IVa (1) I 8 B blaZ, erm(A), sdrM egc Germany, Belgium (1)
76 CC/ST45-t015 IVc (1) I 8 B blaZ, aadD, sdrM egc, sec, sel Germany, Belgium (1)
43, 80 CC/ST59-t316c V (harboring ccrC8) (2) I 8 B blaZ (2), msrA (2), fusB (2), mupA (2), sdrM (2) seb (2), sek (2), seq (2) Australia, Taiwan (1)
83 CC88/ST88-t186 IVa (1) III 8 E blaZ, erm(A), sdrM sec, sel Australia, Japan (1, 57)
70 CC130-t12399 XI (1) III 8 neg sdrM neg Europe, UK (32)
59 CC/ST361-t315c IVg (1) I 8 B blaZ, aacA-aphD, aphA3, sat, tet(K), sdrM, fosB egc Western Australia (32)
84 CC/ST398-t011 IVa (1) I 5 neg blaZ, aacA-aphD, tet(M), sdrM neg Hong Kong, Belgium, Germany (1)
65, 69, 74, 44, 46, 47 CC/ST779-t878 ψSCCmec-SCC-SCCCRISPR (6) III 5 B blaZ (6), aadD (1), fusC (6), mupA (1), sdrM (6) etD (6), edinB (6), seb (1), sed (1), sej (1), ser (1) Uk, Ireland, France, Australia (1, 6)
a

Isolate reference numbers were assigned to each individual sporadic MRSA isolate for inclusion in network trees constructed using complete DNA microarray profile data for all 88 sporadic MRSA isolates using the SplitsTree software package (1, 35) (Fig. 1).

b

IEC types were assigned as described previously. IEC type A includes sea, sak, chp, and scn; IEC type B includes sak, chp, and scn; IEC type C includes chp and scn; IEC type D includes sea, sak, and scn; IEC type E includes sak and scn; IEC type F includes sep, sak, chp, and scn; and IEC type G includes sep, sak, and scn (48). The number of isolates with each IEC type are indicated only when more than one IEC type was identified within a given type combination.

c

MLST was performed on the isolates indicated before the DNA microarray became available. Isolates were selected for MLST based upon spa typing results.

d

Novel MLST sequence types detected.

e

neg, negative.

Fifty-two sporadic MRSA isolates underwent additional PCR-based SCCmec typing to distinguish between SCCmec types and subtypes when the DNA microarray was unable to further differentiate them. This included (i) SCCmec IV subtyping using the method previously described by Milheirico et al. (38), (ii) SCCmec IIA and IIB differentiation using a novel primer, Tn554r (5′-GATAGCAGTATGCCTTAATG-3′), targeting Tn554, which is present only in SCCmec IIA, and a previously described ccrAB2 forward primer (α2) (39), (iii) SCCmec IIIA and IIIB differentiation using a multiplex PCR previously described by Oliveira and de Lencastre (40), and (iv) ccrC allotype identification using a previously described multiplex PCR to differentiate between SCCmec types V (ccrC2) and VT (ccrC2 and ccrC8) (41). In addition, two isolates harboring potentially novel SCCmec types were also further investigated using two previously described multiplex PCR schemes targeting the mec class and the ccr gene complexes (39). Finally, one isolate underwent PCR to confirm the presence of mecC and long-range PCR to confirm the presence of SCCmec XI using previously described primers (2). Long-range PCRs were performed using the Expand Long Template PCR system (Roche Diagnostics GmdH, Lewes, East Sussex, United Kingdom). The following MRSA control reference strains and clinical isolates were used as positive controls for SCCmec typing: AR07.4/0237 (SCCmec IIA/B) (20), E0898 (SCCmec III) (20), CA05 (SCCmec IVa) (42), 8/63-P (SCCmec IVb) (42), JCSC/4788 (SCCmec IVc) (43), JCSC/4469 (SCCmec IVd) (44), M04/0177 (SCCmec IVg) (17), E1749 (SCCmec IVh) (17), WIS (SCCmec V) (45), PM1 (SCCmec VT) (41), and M10/0061 (SCCmec XI) (2).

RESULTS

Genotyping.

Fifty-four different combinations of MLST CC/ST, spa types, and SCCmec types were identified among the 88 isolates, 41 of which were each represented by just one isolate (Table 1). The most prevalent type combination was CC8/ST8-t190-MRSA-IID and SCCM1 (11.4% [10/88]), followed by CC779/ST779-t878-MRSA-ψSCCmec-SCC-SCCCRISPR (6.8% [6/88]) and CC22/ST22-t032-MRSA-IVh (5.7% [5/88]). Three type combinations occurred in 4.5% (4/88) of isolates, including CC8/ST8-t190-MRSA-IIE and SCCM1, CC30/ST36-t018-MRSA-II, and CC45/ST45-t727-MRSA-IVa. Seven type combinations occurred in 2.3% (2/88) of isolates each, including the combination CC1-t127-MRSA-IVa and SCCfus and the combinations ST1-t2279-MRSA-IVa and SCCfus, CC1-t2279-MRSA-IVa, ST5-t045-MRSA-II, ST5-t242-MRSA-VT, CC8/ST8-t008-MRSA-IVa, and ST59-t316-MRSA-V (harboring ccrC8) (Table 1). A total of 37 spa types and 13 MLST-CCs were identified (Table 1). The identification of STs using MLST or the DNA microarray or both was possible for 63/88 isolates, resulting in 17 STs, 4 of which were novel (Table 1). Overall, isolates belonging to CC8 predominated (24/88 [27.3%]), followed by isolates belonging to CC5 (17/88 [19.3%]), CC1 (10/88 [11.4%]), CC22 (9/88 [10.2%]), CC45 (9/88 [10.2%]), CC779 (6/88 [6.8%]), CC30 (6/88 [6.8%]), and CC59 (2/88 [2.3%]). The remaining CCs (6, 78, 398, 361, and 130) were each represented by one isolate (Table 1).

A total of 25 SCCmec types and subtypes were identified, including SCCmec IVa (20.5% [18/88]), which was the most prevalent, followed by SCCmec IID and SCCM1 (11.4% [10/88]), SCCmec II (10.2% [9/88]), SCCmec IVh (10.2% [9/88]), ψSCCmec-SCC-SCCcrispr (6.8% [6/88]), SCCmec IIE and SCCM1, SCCmec VT, and SCCmec IVa and SCCfus (4.5% [4/88]), SCCmec IVc (3.4% [3/88]), and SCCmec IIIB and SCCmec IVg (2.3% [2/88]); six SCCmec types were detected in just one isolate, including SCCmec types IID and III, SCChg IIIA, and SCCmec IVd, VI, and XI (Table 1).

Two isolates (CC5/ST100-t002 and CC45/ST45-t747) carried possible novel SCCmec elements, because mecA was identified, but no ccr gene could be detected by the DNA microarray or by PCR-based SCCmec typing (Table 1). The remaining nine isolates harbored six possible novel SCCmec subtypes (10.2% [9/88]). Of these, three isolates harbored SCCmec elements assigned to previously described SCCmec types, but additional ccr genes were also identified, i.e., SCCmec I and ccrC (ST5-t109), SCCmec II and ccrC (ST36-t018), and SCCmec IV (nonsubtypeable) and ccrB4 (CC5-t067) (Table 1); two isolates (CC5-t002 and ST930-t002) harbored SCCmec IV elements that could not be subtyped (Table 1).

Two novel SCCmec V or VT variants were identified in four isolates due to the carriage of a class C mec complex but unusual combinations of ccr genes. The SCCmec V or VT elements described in the literature to date have been described as harboring class C mec and (i) ccrC allele ccrC1 in SCCmec V (5C) in MRSA strain WIS (45), (ii) ccrC8 and ccrC10 in SCCmec V (5C2&5) in MRSA strain UMCG-M4 (46), or (iii) ccrC2 and ccrC8 in SCCmec VT (5C2&5) in MRSA strain PM1 (41, 47). However, four isolates in the present study carried class C mec, but one harbored ccrC2 only (CC5-t002) and three isolates carried ccrC8 only (CC5-t442 and two ST59-t316) (Table 1).

Overall, SCCmec IV types and subtypes predominated and accounted for 45.5% (40/88) of isolates, followed by SCCmec II (29.5% [26/88]), SCCmec V (9% [8/88]), pseudo element ψSCCmec-SCC-SCCcrispr (6.8% [6/88]), and SCCmec III (3.4% [3/88]) (Table 1). The majority of isolates carried mecA, with just one isolate (CC130-t12399) carrying mecC.

Virulence-associated genes.

Immune evasion cluster (IEC) genes were detected among 84.1% (74/88) of sporadic MRSA isolates and included scn (84% [74/88]), sak (84% [74/88]), chp (44.3% [39/88]), sea (34.1% [30/88]), and sep (6.8% [6/88]) (Table 1). The most common IEC type as defined by Van Wamel et al. was IEC type B (34.1% [30/88]), followed by D (28.4% [25/88]), E (9.1% [8/88]), A (6.8% [6/88]), F (4.5% [4/88]), and G (1.1% [1/88]) (Table 1) (48). Clonal complex 5 exhibited the most IEC types, including IEC types A, B, and D to G, while CC22, CC45, CC59, and CC779 harbored IEC type B only. CC1, CC8, and CC30 harbored multiple IEC types, and all CC8 MRSA isolates harboring SCCmec IID plus SCCM1 and SCCmec IIE plus SCCM1 elements that harbored IEC genes exhibited IEC type D, and this association has been reported previously (Table 1) (21). The accessory gene regulator (agr) allele I was the most dominant agr type (47.7% [42/88]), followed by agr III (27.3% [24/88]), agr II (19.3% [17/88]), and agr IV (6.8% [6/88]) (Table 1). The capsule gene type 5 predominated and was detected in 60.2% (53/88) of isolates (Table 1).

The virulence-associated genes detected among the isolates belonging to the different CCs are shown in Table 1. The most common toxin genes detected were the enterotoxin gene cluster (egc), which was detected in 48.9% (43/88) of sporadic isolates belonging to six CCs, and the enterotoxin A gene sea, which was detected in 34.1% (30/88) belonging to four CCs (1, 5, 8, and 30). The enterotoxin genes sek and seq were harbored by 16% (14/88) of isolates (CC1, -8, and -59), and 11.4% (10/88) of isolates (all CC1) harbored the enterotoxin gene seh. The toxic shock toxin gene tst was detected in 9% (8/88) of isolates, all of which belonged to CC30 (83.3% [/6]) or CC5 (15.8% [3/19]). The enterotoxin genes sec and sel were harbored by 7.9% (7/88) of isolates from three CCs (22, 45, and 78), and seb was detected in 4.5% (4/88) of isolates from three CCs (1, 8, and 59). Various combinations of the enterotoxin genes sed, sej, and ser were detected in 11.4% (10/88) of isolates (CC5, ST30, CC22, and CC779). The ACME-arc genes were detected in 2.3% (2/88) of isolates (both ST8-MRSA-IVa). The exfoliative toxin gene etD and the epidermal cell differentiation inhibitor gene, edinB, were detected among all ST779-MRSA isolates (6.8% [6/88]), and the sep enterotoxin gene was detected in 6.8% (6/88) of isolates in one CC (CC5).

Antimicrobial resistance genes.

The antimicrobial resistance genes detected among the isolates belonging to the different CCs are shown in Table 1. The most prevalent antimicrobial resistance genes detected among the 88 sporadic MRSA isolates other than mec were the beta-lactamase resistance gene blaZ (96.6% [84/88 isolates]) and sdrM, encoding an unspecific efflux pump (89.8% [79/88]). The erm(A) gene (encoding resistance to macrolides, lincosamides, and streptogramin B compounds) was detected in 40.9% of isolates (36/88) belonging to 6/13 CCs (CC1, -5, -8, -30, and -45 and ST88). The aminoglycoside resistance gene aacA-aphD (encoding resistance to amikacin, gentamicin, kanamycin, and tobramycin) was detected in 39.8% of isolates (35/88) in 6/13 CCs (CC5, -8, -22, and -30, ST361, and ST398). Other significant antimicrobial resistance genes detected included the fusidic acid resistance genes fusB and fusC, which were detected in 7.9% (7/88) and 11.4% (11/88) of isolates, respectively. The fusB gene was detected in three CCs (8, 45, and 59), and the fusC gene was detected in two CCs (1 and 779). The mupirocin resistance gene, mupA, was present in 10.2% of isolates (9/88) belonging to seven different CCs (5, 8, 22, 30, 45, 59, and 779).

Two or more resistance genes that encoded resistance to commonly used antimicrobial agents, including aminoglycosides, macrolides-lincosamides, tetracycline, fusidic acid, and mupirocin, were detected among 55.7% (49/88) of isolates and included isolates from all CCs except for the CCs represented by one isolate only (6, 78, 398, 361, and 130) (Table 1).

Similarities between sporadic and global isolates based on microarray data.

Figure 1a shows a graphic representation of the diversity detected among the 88 sporadic MRSA isolates based on DNA microarray profiles. SplitsTree analysis using the transformed microarray profile data separated all 88 isolates into their MLST CCs, and within each CC, similar isolates were grouped closely together. For example, two closely related ACME-arc-positive ST8-MRSA-IVa isolates (isolates 39 and 53) (Table 1) clustered together and were separate from a more distantly related ACME-arc-negative ST8-MRSA-IVd isolate (isolate 35) (Table 1). Figure 1b shows a graphic representation of the relationships between the 88 sporadic MRSA isolates based on array profile data relative to a very large population of global MRSA isolates (n = 3,139). Within the 88 sporadic MRSA isolate population, isolates with specific CCs distributed among global isolates with the same CC in each case (Fig. 1b).

FIG 1.

FIG 1

Network trees generated using the SplitsTree, version 4.11.3, software program (1, 35) and the StaphyType DNA microarray profiles as described previously (1) to visualize the similarities and relationships between CCs and mobile genetic elements, including SCCmec, for the 88 sporadic MRSA isolates investigated and a global population of MRSA isolates. (a) Network tree showing the relationships between the 88 sporadic MRSA isolates investigated. (b) Network tree showing the relationships between the 88 sporadic MRSA isolates investigated in the present study and a previously described global collection of MRSA isolates (n = 3,139) (1). Each sporadic MRSA isolate investigated in the present study is indicated with a number in red (numbers 1 to 88); the details of isolates represented by each number are listed in Table 1. The major CCs identified among isolates in the current study and the previous global study are circled in red, and Roman numerals indicate SCCmec types. In panel b, CCs and STs that were exhibited by MRSA strains from this study are shown in red, and if a CC or ST was not exhibited by any of the 88 sporadic MRSA strains, it is shown in black. MRSA strains from the previously described global population (1) that lacked the Panton-Valentine leukocidin toxin genes lukF/S-PV (pvl) are indicated using black letters on a white background, and pvl-positive MRSA strains are indicated using white letters on a black background. The scale bars show how the length of a branch translates in sequence divergence. The unit is divergent nucleotides divided by the length of the sequence analyzed. Abbreviations: IRR, irregular SCCmec elements; COM, composite or multiple SCCmec elements.

DISCUSSION

Many detailed investigations of the predominant nosocomial MRSA clones prevalent in different regions of the world have been reported, whereas in-depth systematic investigations of sporadically occurring MRSA clones are scarce. The highly clonal ST22-MRSA-IV strain continues to predominate in Irish hospitals, but the prevalence of sporadically occurring MRSA from BSIs increased 2-fold between 2005 and 2011 (19). This study is the first to investigate in detail the molecular epidemiology of sporadic MRSA isolates in Irish hospitals, and it has revealed extensive diversity in genetic backgrounds, SCCmec elements, and virulence and resistance genes. Comparative analysis of DNA microarray data from the 88 sporadic isolates investigated and the corresponding data from 3,139 global MRSA isolates revealed that the relationships between the sporadic MRSA isolates from patients in Irish hospitals reflects the relationships between global MRSA isolates (Fig. 1). An apparently reduced diversity of SCCmec elements in the 88 sporadic MRSA isolates compared to the global MRSA population likely reflects the reduced biodiversity associated with a restricted/insular geographic location (Fig. 1).

A total of 54 MLST, spa, and SCCmec type combinations were identified among the 88 sporadic MRSA isolates investigated, with 49 isolates (55.7%) carrying genes encoding resistance to two or more commonly used antimicrobial agents and 40 (38.6%) harboring two or more virulence-associated genes previously reported to be located on MGEs. Isolates belonging to CC8/ST8-t190-IID/IIE ± SCCM1 predominated. Previous studies have demonstrated a reduced fitness associated with larger SCCmec elements (49, 50), and we previously speculated that the potential fitness cost associated with carrying a large SCCmec-SCC composite island (CI) may have contributed to the decline of ST8-MRSA-IIA-IIE and SCCM1 (21) in Irish hospitals. However, the many resistance genes detected among isolates of this clone may also have contributed to its decline due to reduced fitness, but other lineage-specific factors may also have contributed (51).

The second most prevalent clone, ST779-t878-MRSA-ψSCCmec-SCC-SCCcrispr, also carried multiple resistance genes and a large SCC-CI element that may have originated in coagulase-negative staphylococci (CoNS). We recently reported the emergence of this clone in Irish hospitals (6), and ST779-MRSA has also been reported sporadically in Australia, Canada, Germany, Malaysia, Thailand, the United Arab Emirates, and the United Kingdom (http://saureus.mlst.net/) (2, 6, 24, 52). Fitness costs associated with the carriage of a large SCC-CI and multiple MGE-located resistance genes may curtail the widespread emergence of this clone in the absence of selective pressure.

Extensive diversity was detected among the SCCmec elements harbored by the sporadic MRSA isolates and included 25 different SCCmec types and subtypes encompassing types and/or subtypes of SCCmec types I to VI, SCCmec type XI, two possible novel SCCmec elements, and six possible novel SCCmec subtypes. SCCmec type IV predominated, accounting for almost half of the isolates. Since SCCmec IV is also the SCCmec type of the ST22-MRSA-IV clone endemic in Irish hospitals for the last decade (17), it is clear that SCCmec IV is the dominant SCCmec element among all nosocomial MRSA isolates in Ireland. However, eight different subtypes were identified among the sporadic isolates, with SCCmec IVa being the most common (20.5% [18/40]). In contrast, SCCmec IVh predominates among isolates of the endemic ST22-MRSA-IV clone (17). It is not possible to discriminate between most SCCmec IV subtypes using the DNA microarray, and considering that these are associated with particular pandemic MRSA clones, e.g., SCCmec IVh in ST22 and SCCmec IVa in ST8/USA300, it is essential that detailed SCCmec IV subtyping be performed to ensure effective tracking and typing of these clones.

SCCmec V and VT subtyping identified novel SCCmec V subtypes and provided further evidence of the diversity present in SCCmec V elements (46), including ccrC alleles. The CC59-MRSA-V clone usually harbors two ccrC1 complexes (ccrC1 allele 2 and ccrC allele 8) (47). However, the two CC59-MRSA-V isolates and a CC5-t442-MRSA isolate identified in the study harbored only ccrC1 allele 8. Additionally, a CC5-t002-MRSA isolate harbored an SCCmec V element with just one ccrC allotype, ccrC2. These may represent possible SCCmec V variants or precursors in two separate CCs, CC5 and CC59.

The majority of isolates investigated had genotypes generally considered to be health care associated, including ST8-MRSA-IID/IIE ± SCCM1, ST239-MRSA-III, ST36-MRSA-II, ST22-MRSA-IV, ST45-MRSA-IV, ST5-MRSA-II, and ST361-t315-MRSA-IVg (1, 53), each of which, apart from the ST361 MRSA-IVg isolate, was previously identified in Ireland as either predominant or sporadic strains (2, 20). Many of these clones predominate or have predominated in hospitals in other countries, and no major differences were noted between these isolates and those reported previously (1). A number of isolates with CC/ST and SCCmec type combinations commonly associated with pvl-positive community-associated MRSA (CA-MRSA) clones were also detected, including CC1-MRSA-IV (1, 54), CC59-MRSA-V (47, 55), ST8-t008-MRSA-IVa (1), CC5-MRSA-IV (1), CC5-MRSA-V (56), and CC88/ST88-t186-MRSA-IVa (1, 57). It should be noted that potential CA-MRSA-associated clones may be underrepresented in the present study due to the exclusion of pvl-positive sporadic MRSA isolates. The prevalence of CA-MRSA (both pvl positive and negative) among patients in Irish hospitals remains to be determined.

This study also found further evidence of the possible zoonotic spread of MRSA in Ireland. First, a CC130-MRSA-XI isolate recovered in 2007 was identified. We previously reported the recovery in 2010 of two sporadic CC130-MRSA-XI isolates from separate hospitals (2). The newly identified isolate exhibited a previously unreported spa type (t12399) harboring two additional spa repeats compared to spa type t843 exhibited by the CC130 MRSA isolates recovered in 2010 (2). The isolate was recovered from an elderly patient in the community who had previously been hospitalized on several occasions and who lived adjacent to a farm. Since its first detection, SCCmec XI has been reported sporadically among MRSA isolates belonging to a number of animal-associated MRSA lineages (predominantly CC130) in many different European countries from human and animal sources (32), and several studies have provided evidence for the zoonotic spread of these strains (58, 59). Other clones of possible animal origin were also identified, all recovered between 2007 and 2011, including the equine-associated ST8-t064-MRSA-IVa clone (60, 61), as well as the livestock-associated clone ST398-t011-MRSA-IVa and CC6-MRSA-IVh, which has been linked with camels (1, 62, 63). These findings highlight the importance of animals as a reservoir for MRSA and for effective surveillance to minimize the spread of these clones in hospitals.

The prevalence and diversity of resistance and virulence genes identified among the sporadic MRSA isolates also highlight the extensive reservoir of these genes that exist within the population of Irish MRSA. This, coupled with the range of genetic backgrounds of the isolates, highlights the potential for spread of these resistance genes and thus our ability to treat MRSA colonization and infection. For example, a high rate of the carriage of macrolide (57.9%) and aminoglycoside (43.4%) resistance genes was observed among isolates belonging to an extensive range of genetic backgrounds. Additionally, the high-level mupirocin resistance gene mupA, known to be carried on conjugative plasmids (64), was identified in 9/88 (10.2%) isolates belonging to seven different genetic backgrounds. Mupirocin is commonly used for MRSA nasal decolonization, and previous reports from Ireland have reported high-level mupirocin resistance rates among MRSA isolates from BSIs, ranging from 1.4% between 1999 and 2005 to 3.1% in 2011, predominantly among ST22-MRSA-IV and ST8-MRSA-IIA-IIE isolates (19). Lastly, in Ireland the rate of phenotypic fusidic acid resistance among MRSA isolates from BSIs increased from <10% to 34% between 1999 and 2011 (19). In the present study, 18/88 (20.5%) sporadic isolates harbored either the plasmid-located fusB gene or the SCC-associated fusC gene. More stringent use of these antimicrobial agents is warranted so that resistance does not become more widespread.

Few studies focused primarily on the detailed characterization of sporadic MRSA isolates. The main emphasis of most studies that reported such isolates concentrated on identifying the main MRSA lineages present in large populations of MRSA isolates from particular countries or from several hospitals (20, 65, 66). For example, while reporting the clonal replacement of CC5/ST228-MRSA-I and CC5-MRSA-II by the emerging CC22-MRSA-IV and CC45-MRSA-IV clones as the predominant nosocomial strains over an 11-year period in a German tertiary care hospital, Albrecht et al. also identified 17 pvl-negative sporadic MRSA isolates among 778 isolates investigated, including CC7-MRSA-IV, CC97-MRSA-IV, CC88-MRSA-IV, and CC30/ST36-MRSA-II (67), the former two of which were also identified in the present study. The diversity identified among the Irish sporadic MRSA isolates investigated in this study spans most of the lineages seen at the global level (Fig. 1). This may be because the strains, or at least some of them, have at some stage been endemic in Ireland since their evolutionary origin. However, it is important to emphasize that the origin of some MRSA strains can be polyphyletic resulting from multiple transmissions of identical or similar SCCmec elements from MRSA or CoNS into methicillin-susceptible S. aureus (MSSA) of one clonal lineage (1). Recurrent importation of MRSA strains from other countries is also likely to have been another significant factor contributing to the diversity found among the sporadic MRSA isolates. The latter suggestion is reflected by the findings of a recent study from this laboratory on pvl-positive MRSA recovered in Ireland over the last decade that revealed frequent importation of MRSA strains, particularly in recent years (28). While the increasing prevalence of sporadic MRSA strains in Ireland may be due to an increase in their importation or to the local emergence of strains, the decreasing prevalence of ST22-MRSA-IV in Irish hospitals may also have contributed, allowing for the emergence of these sporadic MRSA with enhanced virulence and resistance potential. However, further studies of both sporadic and endemic MRSA as well as MSSA are required in order to determine this.

In conclusion, the diversity detected among the 88 representative sporadic MRSA isolates, including SCCmec and SCC-associated elements and virulence-associated and antimicrobial resistance genes, and the number of different genetic lineages identified by MLST, spa typing, and DNA microarray analysis provide further evidence of the need for effective surveillance of this genetically diverse reservoir. Exchange of genetic material between these and other more prevalent MRSA strains may contribute to the emergence of successful MRSA strains in the future. Shore et al. previously demonstrated that there is a history of strain replacement approximately once per decade in Ireland, and therefore, it is important that emerging MRSA strains be detected early (20). The ST22-MRSA-IV clone has predominated for more than a decade in Irish hospitals, and its recent decline in prevalence suggests that a novel strain(s) may emerge in the near future.

Supplementary Material

Supplemental material

ACKNOWLEDGMENTS

This work was supported by the Microbiology Research Unit, Dublin Dental University Hospital.

We thank the staff of the Irish National MRSA Reference Laboratory, past and present, in particular Angela Rossney, for their ongoing support and collaboration in investigating large collections of MRSA and MSSA isolates. MRSA control isolates were kindly provided by Teuryo Ito, Juntendo University, Japan, and Herminia de Lencastre, Rockfeller University, New York.

Footnotes

Published ahead of print 6 January 2014

Supplemental material for this article may be found at http://dx.doi.org/10.1128/AAC.02653-13.

REFERENCES

  • 1.Monecke S, Coombs G, Shore AC, Coleman DC, Akpaka P, Borg M, Chow H, Ip M, Jatzwauk L, Jonas D, Kadlec K, Kearns A, Laurent F, O'Brien FG, Pearson J, Ruppelt A, Schwarz S, Scicluna E, Slickers P, Tan HL, Weber S, Ehricht R. 2011. A field guide to pandemic, epidemic and sporadic clones of methicillin-resistant Staphylococcus aureus. PLoS One 6:e17936. 10.1371/journal.pone.0017936 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Shore AC, Deasy EC, Slickers P, Brennan G, O'Connell B, Monecke S, Ehricht R, Coleman DC. 2011. Detection of staphylococcal cassette chromosome mec type XI carrying highly divergent mecA, mecI, mecR1, blaZ, and ccr genes in human clinical isolates of clonal complex 130 methicillin-resistant Staphylococcus aureus. Antimicrob. Agents Chemother. 55:3765–3773. 10.1128/AAC.00187-11 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Ito T, Hiramatsu K, Tomasz A, de Lencastre H, Perreten V, Holden MT, Coleman DC, Goering R, Giffard PM, Skov RL, Zhang K, Westh H, O'Brien F, Tenover FC, Oliveira DC, Boyle-Vavra S, Laurent F, Kearns AM, Kreiswirth B, Ko KS, Grundmann H, Sollid JE, John JF, Jr, Daum R, Soderquist B, Buist G. 2012. Guidelines for reporting novel mecA gene homologues. Antimicrob. Agents Chemother. 56:4997–4999. 10.1128/AAC.01199-12 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Kim C, Milheirico C, Gardete S, Holmes MA, Holden MT, de Lencastre H, Tomasz A. 2012. Properties of a novel PBP2A protein homolog from Staphylococcus aureus strain LGA251 and its contribution to the beta-lactam-resistant phenotype. J. Biol. Chem. 287:36854–36863. 10.1074/jbc.M112.395962 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Shore AC, Coleman DC. 2013. Staphylococcal cassette chromosome mec: recent advances and new insights. Int. J. Med. Microbiol. 303:350–359. 10.1016/j.ijmm.2013.02.002 [DOI] [PubMed] [Google Scholar]
  • 6.Kinnevey PM, Shore AC, Brennan GI, Sullivan DJ, Ehricht R, Monecke S, Slickers P, Coleman DC. 2013. Emergence of sequence type 779 methicillin-resistant Staphylococcus aureus (MRSA) harboring a novel pseudo staphylococcal cassette chromosome mec (SCCmec)-SCC-SCCCRISPR composite element in Irish hospitals. Antimicrob. Agents Chemother. 57:524–531 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.International Working Group on the Classification of Staphylococcal Cassette Chromosome Elements (IWG-SCC). 2009. Classification of staphylococcal cassette chromosome mec (SCCmec): guidelines for reporting novel SCCmec elements. Antimicrob. Agents Chemother. 53:4961–4967. 10.1128/AAC.00579-09 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Malachowa N, DeLeo FR. 2010. Mobile genetic elements of Staphylococcus aureus. Cell. Mol. Life Sci. 67:3057–3071. 10.1007/s00018-010-0389-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Chambers HF, Deleo FR. 2009. Waves of resistance: Staphylococcus aureus in the antibiotic era. Nat. Rev. Microbiol. 7:629–641. 10.1038/nrmicro2200 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Jevons MP. 1961. Celbanin-resistant staphylococci. Br. Med. J. i:124–125 [Google Scholar]
  • 11.Hone R, Keane CT. 1974. Characteristics of methicillin resistant Staphylococcus aureus. Irish J. Med. Sci. 143:145–154. 10.1007/BF03004756 [DOI] [PubMed] [Google Scholar]
  • 12.Coleman DC, Pomeroy H, Estridge JK, Keane CT, Cafferkey MT, Hone R, Foster TJ. 1985. Susceptibility to antimicrobial agents and analysis of plasmids in gentamicin- and methicillin-resistant Staphylococcus aureus from Dublin hospitals. J. Med. Microbiol. 20:157–167. 10.1099/00222615-20-2-157 [DOI] [PubMed] [Google Scholar]
  • 13.Carroll JD, Pomeroy HM, Russell RJ, Arbuthnott JP, Keane CT, McCormick OM, Coleman DC. 1989. A new methicillin- and gentamicin-resistant Staphylococcus aureus in Dublin: molecular genetic analysis. J. Med. Microbiol. 28:15–23. 10.1099/00222615-28-1-15 [DOI] [PubMed] [Google Scholar]
  • 14.Rossney AS, Keane CT. 2002. Strain variation in the MRSA population over a 10-year period in one Dublin hospital. Eur. J. Clin. Microbiol. Infect. Dis. 21:123–126. 10.1007/s10096-001-0677-x [DOI] [PubMed] [Google Scholar]
  • 15.Rossney AS, McDonald P, Humphreys H, Glynn GM, Keane CT. 2003. Antimicrobial resistance and epidemiological typing of methicillin-resistant Staphylococcus aureus in Ireland (North and South), 1999. Eur. J. Clin. Microbiol. Infect. Dis. 22:379–381. 10.1007/s10096-003-0917-3 [DOI] [PubMed] [Google Scholar]
  • 16.Rossney AS, Lawrence MJ, Morgan PM, Fitzgibbon MM, Shore A, Coleman DC, Keane CT, O'Connell B. 2006. Epidemiological typing of MRSA isolates from blood cultures taken in Irish hospitals participating in the European Antimicrobial Resistance Surveillance System (1999–2003). Eur. J. Clin. Microbiol. Infect. Dis. 25:79–89. 10.1007/s10096-006-0091-5 [DOI] [PubMed] [Google Scholar]
  • 17.Shore AC, Rossney AS, Kinnevey PM, Brennan OM, Creamer E, Sherlock O, Dolan A, Cunney R, Sullivan DJ, Goering RV, Humphreys H, Coleman DC. 2010. Enhanced discrimination of highly clonal ST22-methicillin-resistant Staphylococcus aureus IV isolates achieved by combining spa, dru, and pulsed-field gel electrophoresis typing data. J. Clin. Mcrobiol. 48:1839–1852. 10.1128/JCM.02155-09 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.EARS-Net. 2012. EARS-Net MRSA summary report. National MRSA Reference Laboratory, Dublin, Ireland: http://www.stjames.ie/Departments/DepartmentsA-Z/N/NationalMRSAReferenceLaboratory/DepartmentinDepth/MRSAQ412.pdf [Google Scholar]
  • 19.NMRSARL. 2011. National MRSA Reference Laboratory annual report. National MRSA Reference Laboratory, Dublin, Ireland: http://www.stjames.ie/Departments/DepartmentsA-Z/N/NationalMRSAReferenceLaboratory/DepartmentinDepth/NMRSARL%20Annual%20Report%202011.pdf Accessed 9 November 2011 [Google Scholar]
  • 20.Shore A, Rossney AS, Keane CT, Enright MC, Coleman DC. 2005. Seven novel variants of the staphylococcal chromosomal cassette mec in methicillin-resistant Staphylococcus aureus isolates from Ireland. Antimicrob. Agents Chemother. 49:2070–2083. 10.1128/AAC.49.5.2070-2083.2005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Shore AC, Brennan OM, Deasy EC, Rossney AS, Kinnevey PM, Ehricht R, Monecke S, Coleman DC. 2012. DNA microarray profiling of a diverse collection of nosocomial methicillin-resistant Staphylococcus aureus isolates assigns the majority to the correct sequence type and staphylococcal cassette chromosome mec (SCCmec) type and results in the subsequent identification and characterization of novel SCCmec-SCCM1 composite islands. Antimicrob. Agents Chemother. 56:5340–5355. 10.1128/AAC.01247-12 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Dominguez MA, de Lencastre H, Linares J, Tomasz A. 1994. Spread and maintenance of a dominant methicillin-resistant Staphylococcus aureus (MRSA) clone during an outbreak of MRSA disease in a Spanish hospital. J. Clin. Microbiol. 32:2081–2087 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Oliveira D, Santos-Sanches I, Mato R, Tamayo M, Ribeiro G, Costa D, de Lencastre H. 1998. Virtually all methicillin-resistant Staphylococcus aureus (MRSA) infections in the largest Portuguese teaching hospital are caused by two internationally spread multiresistant strains: the ‘Iberian' and the ‘Brazilian' clones of MRSA. Clin. Microbiol. Infect. 4:373–384. 10.1111/j.1469-0691.1998.tb00081.x [DOI] [PubMed] [Google Scholar]
  • 24.Lim KT, Yeo CC, Suhaili Z, Thong KL. 2012. Comparison of methicillin-resistant and methicillin-sensitive Staphylococcus aureus strains isolated from a tertiary hospital in Terengganu, Malaysia. Jpn. J. Infect. Dis. 65:502–509. 10.7883/yoken.65.502 [DOI] [PubMed] [Google Scholar]
  • 25.Mohamed DH, Saberesheikh S, Kearns AM, Saunders NA. 2012. Putative link between Staphylococcus aureus bacteriophage serotype and community association. Int. J. Med. Microbiol. 302:135–144. 10.1016/j.ijmm.2012.02.002 [DOI] [PubMed] [Google Scholar]
  • 26.Kuhn G, Koessler T, Melles DC, Francois P, Huyghe A, Dunman P, Vos MC, Zanetti G, Schrenzel J, van Belkum A, Blanc DS. 2010. Comparative genomics of epidemic versus sporadic Staphylococcus aureus strains does not reveal molecular markers for epidemicity. Infect. Genet. Evol. 10:89–96. 10.1016/j.meegid.2009.10.011 [DOI] [PubMed] [Google Scholar]
  • 27.Melin S, Melin S, Haeggman S, Melin S, Haeggman S, Olsson-Liljequist B, Sjolund M, Nilsson PA, Isaksson B, Lofgren S, Matussek A. 2009. Epidemiological typing of methicillin-resistant Staphylococcus aureus (MRSA): spa typing versus pulsed-field gel electrophoresis. Scand. J. Infect. Dis. 41:433–439. 10.1080/00365540902962749 [DOI] [PubMed] [Google Scholar]
  • 28.Shore AC, Tecklenborg SC, Brennan GI, Ehricht R, Monecke S, Coleman DC. 2014. Panton-Valentine leukocidin-positive Staphylococcus aureus in Ireland from 2002 to 2011: 21 clones, frequent importation of clones, temporal shifts of predominant methicillin-resistant S. aureus isolates, and increasing multiresistance. J. Clin. Microbiol. 52:859–870. 10.1128/JCM.02799-13 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Tenover FC, Arbeit RD, Goering RV, Mickelsen PA, Murray BE, Persing DH, Swaminathan B. 1995. Interpreting chromosomal DNA restriction patterns produced by pulsed-field gel electrophoresis: criteria for bacterial strain typing. J. Clin. Microbiol. 33:2233–2239 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Rossney AS, Shore AC, Morgan PM, Fitzgibbon MM, O'Connell B, Coleman DC. 2007. The emergence and importation of diverse genotypes of methicillin-resistant Staphylococcus aureus (MRSA) harboring the Panton-Valentine leukocidin gene (pvl) reveal that pvl is a poor marker for community-acquired MRSA strains in Ireland. J. Clin. Microbiol. 45:2554–2563. 10.1128/JCM.00245-07 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Shore AC, Rossney AS, Brennan OM, Kinnevey PM, Humphreys H, Sullivan DJ, Goering RV, Ehricht R, Monecke S, Coleman DC. 2011. Characterization of a novel arginine catabolic mobile element (ACME) and staphylococcal chromosomal cassette mec composite island with significant homology to Staphylococcus epidermidis ACME type II in methicillin-resistant Staphylococcus aureus genotype ST22-MRSA-IV. Antimicrob. Agents Chemother. 55:1896–1905. 10.1128/AAC.01756-10 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Monecke S, Gavier-Widen D, Mattsson R, Rangstrup-Christensen L, Lazaris A, Coleman DC, Shore AC, Ehricht R. 2013. Detection of mecC-positive Staphylococcus aureus (CC130-MRSA-XI) in diseased European hedgehogs (Erinaceus europaeus) in Sweden. PLoS One 8:e66166. 10.1371/journal.pone.0066166 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Monecke S, Slickers P, Ehricht R. 2008. Assignment of Staphylococcus aureus isolates to clonal complexes based on microarray analysis and pattern recognition. FEMS Immunol. Med. Microbiol. 53:237–251. 10.1111/j.1574-695X.2008.00426.x [DOI] [PubMed] [Google Scholar]
  • 34.Monecke S, Jatzwauk L, Weber S, Slickers P, Ehricht R. 2008. DNA microarray-based genotyping of methicillin-resistant Staphylococcus aureus strains from Eastern Saxony. Clin. Microbiol. Infect. 14:534–545. 10.1111/j.1469-0691.2008.01986.x [DOI] [PubMed] [Google Scholar]
  • 35.Huson DH, Bryant D. 2006. Application of phylogenetic networks in evolutionary studies. Mol. Biol. Evol. 23:254–267. 10.1093/molbev/msj030 [DOI] [PubMed] [Google Scholar]
  • 36.Mellmann A, Weniger T, Berssenbrugge C, Rothganger J, Sammeth M, Stoye J, Harmsen D. 2007. Based Upon Repeat Pattern (BURP): an algorithm to characterize the long-term evolution of Staphylococcus aureus populations based on spa polymorphisms. BMC Microbiol. 7:98. 10.1186/1471-2180-7-98 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Enright MC, Day NP, Davies CE, Peacock SJ, Spratt BG. 2000. Multilocus sequence typing for characterization of methicillin-resistant and methicillin-susceptible clones of Staphylococcus aureus. J. Clin. Microbiol. 38:1008–1015 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Milheirico C, Oliveira DC, de Lencastre H. 2007. Multiplex PCR strategy for subtyping the staphylococcal cassette chromosome mec type IV in methicillin-resistant Staphylococcus aureus: ‘SCCmec IV multiplex.' J. Antimicrob. Chemother. 60:42–48. 10.1093/jac/dkm112 [DOI] [PubMed] [Google Scholar]
  • 39.Kondo Y, Ito T, Ma XX, Watanabe S, Kreiswirth BN, Etienne J, Hiramatsu K. 2007. Combination of multiplex PCRs for staphylococcal cassette chromosome mec type assignment: rapid identification system for mec, ccr, and major differences in junkyard regions. Antimicrob. Agents Chemother. 51:264–274. 10.1128/AAC.00165-06 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Oliveira DC, de Lencastre H. 2002. Multiplex PCR strategy for rapid identification of structural types and variants of the mec element in methicillin-resistant Staphylococcus aureus. Antimicrob. Agents Chemother. 46:2155–2161. 10.1128/AAC.46.7.2155-2161.2002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Higuchi W, Takano T, Teng LJ, Yamamoto T. 2008. Structure and specific detection of staphylococcal cassette chromosome mec type VII. Biochem. Biophys. Res. Commun. 377:752–756. 10.1016/j.bbrc.2008.10.009 [DOI] [PubMed] [Google Scholar]
  • 42.Ma XX, Ito T, Tiensasitorn C, Jamklang M, Chongtrakool P, Boyle-Vavra S, Daum RS, Hiramatsu K. 2002. Novel type of staphylococcal cassette chromosome mec identified in community-acquired methicillin-resistant Staphylococcus aureus strains. Antimicrob. Agents Chemother. 46:1147–1152. 10.1128/AAC.46.4.1147-1152.2002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Kishii K, Ito T, Watanabe S, Okuzumi K, Hiramatsu K. 2008. Recurrence of heterogeneous methicillin-resistant Staphylococcus aureus (MRSA) among the MRSA clinical isolates in a Japanese university hospital. J. Antimicrob. Chemother. 62:324–328. 10.1093/jac/dkn186 [DOI] [PubMed] [Google Scholar]
  • 44.Ma XX, Ito T, Chongtrakool P, Hiramatsu K. 2006. Predominance of clones carrying Panton-Valentine leukocidin genes among methicillin-resistant Staphylococcus aureus strains isolated in Japanese hospitals from 1979 to 1985. J. Clin. Microbiol. 44:4515–4527. 10.1128/JCM.00985-06 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Ito T, Ma XX, Takeuchi F, Okuma K, Yuzawa H, Hiramatsu K. 2004. Novel type V staphylococcal cassette chromosome mec driven by a novel cassette chromosome recombinase, ccrC. Antimicrob. Agents Chemother. 48:2637–2651. 10.1128/AAC.48.7.2637-2651.2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Chlebowicz MA, Nganou K, Kozytska S, Arends JP, Engelmann S, Grundmann H, Ohlsen K, van Dijl JM, Buist G. 2010. Recombination between ccrC genes in a type V (5C2&5) staphylococcal cassette chromosome mec (SCCmec) of Staphylococcus aureus ST398 leads to conversion from methicillin resistance to methicillin susceptibility in vivo. Antimicrob. Agents Chemother. 54:783–791. 10.1128/AAC.00696-09 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Coombs GW, Monecke S, Ehricht R, Slickers P, Pearson JC, Tan HL, Christiansen KJ, O'Brien FG. 2010. Differentiation of clonal complex 59 community-associated methicillin-resistant Staphylococcus aureus in Western Australia. Antimicrob. Agents Chemother. 54:1914–1921. 10.1128/AAC.01287-09 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Van Wamel WJ, Rooijakkers SH, Ruyken M, van Kessel KP, van Strijp JA. 2006. The innate immune modulators staphylococcal complement inhibitor and chemotaxis inhibitory protein of Staphylococcus aureus are located on beta-hemolysin-converting bacteriophages. J. Bacteriol. 188:1310–1315. 10.1128/JB.188.4.1310-1315.2006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Ender M, McCallum N, Adhikari R, Berger-Bachi B. 2004. Fitness cost of SCCmec and methicillin resistance levels in Staphylococcus aureus. Antimicrob. Agents Chemother. 48:2295–2297. 10.1128/AAC.48.6.2295-2297.2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Knight GM, Budd EL, Lindsay JA. 2013. Large mobile genetic elements carrying resistance genes that do not confer a fitness burden in healthcare-associated methicillin-resistant Staphylococcus aureus. Microbiology 159:1661–1672. 10.1099/mic.0.068551-0 [DOI] [PubMed] [Google Scholar]
  • 51.Nielsen KL, Pedersen TM, Udekwu KI, Petersen A, Skov RL, Hansen LH, Hughes D, Frimodt-Moller N. 2012. Fitness cost: a bacteriological explanation for the demise of the first international methicillin-resistant Staphylococcus aureus epidemic. J. Antimicrob. Chemother. 67:1325–1332. 10.1093/jac/dks051 [DOI] [PubMed] [Google Scholar]
  • 52.Coombes G, Pearson J, Christiansen K, Nimmo G. 2010. Staphylococcus aureus programme 2010 (SAP 2010). Community survey. MRSA epidemiology and typing report. The Australian Group for Antimicrobial Resistance. http://www.agargroup.org/files/FED%20REPORT%20SAP210%20MRSA%20FINAL%20shrink.pdf
  • 53.Afroz S, Kobayashi N, Nagashima S, Alam MM, Hossain AB, Rahman MA, Islam MR, Lutfor AB, Muazzam N, Khan MA, Paul SK, Shamsuzzaman AK, Mahmud MC, Musa AK, Hossain MA. 2008. Genetic characterization of Staphylococcus aureus isolates carrying Panton-Valentine leukocidin genes in Bangladesh. Jpn. J. Infect. Dis. 61:393–396 [PubMed] [Google Scholar]
  • 54.Mediavilla JR, Chen L, Mathema B, Kreiswirth BN. 2012. Global epidemiology of community-associated methicillin resistant Staphylococcus aureus (CA-MRSA). Curr. Opin. Microbiol. 15:588–595. 10.1016/j.mib.2012.08.003 [DOI] [PubMed] [Google Scholar]
  • 55.Rolo J, Miragaia M, Turlej-Rogacka A, Empel J, Bouchami O, Faria NA, Tavares A, Hryniewicz W, Fluit AC, de Lencastre H. 2012. High genetic diversity among community-associated Staphylococcus aureus in Europe: results from a multicenter study. PLoS One 7:e34768. 10.1371/journal.pone.0034768 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Espadinha D, Faria NA, Miragaia M, Lito LM, Melo-Cristino J, de Lencastre H. 2013. Extensive dissemination of methicillin-resistant Staphylococcus aureus (MRSA) between the hospital and the community in a country with a high prevalence of nosocomial MRSA. PLoS One 8:e59960. 10.1371/journal.pone.0059960 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Hisata K, Kuwahara-Arai K, Yamanoto M, Ito T, Nakatomi Y, Cui L, Baba T, Terasawa M, Sotozono C, Kinoshita S, Yamashiro Y, Hiramatsu K. 2005. Dissemination of methicillin-resistant staphylococci among healthy Japanese children. J. Clin. Microbiol. 43:3364–3372. 10.1128/JCM.43.7.3364-3372.2005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Petersen A, Stegger M, Heltberg O, Christensen J, Zeuthen A, Knudsen LK, Urth T, Sorum M, Schouls L, Larsen J, Skov R, Larsen AR. 2013. Epidemiology of methicillin-resistant Staphylococcus aureus carrying the novel mecC gene in Denmark corroborates a zoonotic reservoir with transmission to humans. Clin. Microbiol. Infect. 19:E16–E22. 10.1111/1469-0691.12036 [DOI] [PubMed] [Google Scholar]
  • 59.Harrison EM, Paterson GK, Holden MT, Larsen J, Stegger M, Larsen AR, Petersen A, Skov RL, Christensen JM, Bak Zeuthen A, Heltberg O, Harris SR, Zadoks RN, Parkhill J, Peacock SJ, Holmes MA. 2013. Whole genome sequencing identifies zoonotic transmission of MRSA isolates with the novel mecA homologue mecC. EMBO Mol. Med. 5:509–515. 10.1002/emmm.201202413 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Abbott Y, Leonard FC, Markey BK. 2010. Detection of three distinct genetic lineages in methicillin-resistant Staphylococcus aureus (MRSA) isolates from animals and veterinary personnel. Epidemiol. Infect. 138:764–771. 10.1017/S0950268809991580 [DOI] [PubMed] [Google Scholar]
  • 61.van Duijkeren E, Moleman M, Sloet van Oldruitenborgh-Oosterbaan MM, Multem J, Troelstra A, Fluit AC, van Wamel WJ, Houwers DJ, de Neeling AJ, Wagenaar JA. 2010. Methicillin-resistant Staphylococcus aureus in horses and horse personnel: an investigation of several outbreaks. Vet. Microbiol. 141:96–102. 10.1016/j.vetmic.2009.08.009 [DOI] [PubMed] [Google Scholar]
  • 62.Monecke S, Skakni L, Hasan R, Ruppelt A, Ghazal SS, Hakawi A, Slickers P, Ehricht R. 2012. Characterisation of MRSA strains isolated from patients in a hospital in Riyadh, Kingdom of Saudi Arabia. BMC Microbiol. 12:146. 10.1186/1471-2180-12-146 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Brennan GI, O'Connell B, Coleman DC, Shore AC. 2012. First Irish report of livestock-associated MRSA strain. Epi-Insight 13(10). http://ndsc.newsweaver.ie/epiinsight/1c8fwftl674?a=1&p=28075745& t=17517774 [Google Scholar]
  • 64.McDougal LK, Fosheim GE, Nicholson A, Bulens SN, Limbago BM, Shearer JE, Summers AO, Patel JB. 2010. Emergence of resistance among USA300 methicillin-resistant Staphylococcus aureus isolates causing invasive disease in the United States. Antimicrob. Agents Chemother. 54:3804–3811. 10.1128/AAC.00351-10 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Xiao M, Wang H, Zhao Y, Mao LL, Brown M, Yu YS, O'Sullivan MV, Kong F, Xu YC. 2013. National surveillance of methicillin-resistant Staphylococcus aureus (MRSA) in China highlights a still evolving epidemiology with fifteen novel emerging multilocus sequence types. J. Clin. Microbiol. 51:3638–3644. 10.1128/JCM.01375-13 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Collery MM, Smyth DS, Twohig JM, Shore AC, Coleman DC, Smyth CJ. 2008. Molecular typing of nasal carriage isolates of Staphylococcus aureus from an Irish university student population based on toxin gene PCR, agr locus types and multiple locus, variable number tandem repeat analysis. J. Med. Microbiol. 57:348–358. 10.1099/jmm.0.47734-0 [DOI] [PubMed] [Google Scholar]
  • 67.Albrecht N, Jatzwauk L, Slickers P, Ehricht R, Monecke S. 2011. Clonal replacement of epidemic methicillin-resistant Staphylococcus aureus strains in a German university hospital over a period of eleven years. PLoS One 6:e28189. 10.1371/journal.pone.0028189 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Scicluna EA, Shore AC, Thurmer A, Ehricht R, Slickers P, Borg MA, Coleman DC, Monecke S. 2010. Characterisation of MRSA from Malta and the description of a Maltese epidemic MRSA strain. Eur. J. Clin. Microbiol. Infect. Dis. 29:163–170. 10.1007/s10096-009-0834-1 [DOI] [PubMed] [Google Scholar]
  • 69.Hudson LO, Reynolds C, Spratt BG, Enright MC, Quan V, Kim D, Hannah P, Mikhail L, Alexander R, Moore DF, Godoy D, Bishop CJ, Huang SS. 2013. Diversity of methicillin-resistant Staphylococcus aureus strains isolated from residents of 26 nursing homes in Orange County, California. J. Clin. Microbiol. 51:3788–3795. 10.1128/JCM.01708-13 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Lozano C, Porres-Osante N, Crettaz J, Rojo-Bezares B, Benito D, Olarte I, Zarazaga M, Saenz Y, Torres C. 2013. Changes in genetic lineages, resistance, and virulence in clinical methicillin-resistant Staphylococcus aureus in a Spanish hospital. J. Infect. Chemother. 19:233–242. 10.1007/s10156-012-0486-4 [DOI] [PubMed] [Google Scholar]
  • 71.Coombs GW, Monecke S, Pearson JC, Tan HL, Chew YK, Wilson L, Ehricht R, O'Brien FG, Christiansen KJ. 2011. Evolution and diversity of community-associated methicillin-resistant Staphylococcus aureus in a geographical region. BMC Microbiol. 11:215. 10.1186/1471-2180-11-215 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Morales G, Picazo JJ, Baos E, Candel FJ, Arribi A, Pelaez B, Andrade R, de la Torre MA, Fereres J, Sanchez-Garcia M. 2010. Resistance to linezolid is mediated by the cfr gene in the first report of an outbreak of linezolid-resistant Staphylococcus aureus. Clin. Infect. Dis. 50:821–825. 10.1086/650574 [DOI] [PubMed] [Google Scholar]
  • 73.Huber H, Koller S, Giezendanner N, Stephan R, Zweifel C. 2010. Prevalence and characteristics of meticillin-resistant Staphylococcus aureus in humans in contact with farm animals, in livestock, and in food of animal origin, Switzerland, 2009. Eur. Surveill. 15(16):19542 http://www.eurosurveillance.org/ViewArticle.aspx?ArticleId=19542 [PubMed] [Google Scholar]
  • 74.Plano LR, Shibata T, Garza AC, Kish J, Fleisher JM, Sinigalliano CD, Gidley ML, Withum K, Elmir SM, Hower S, Jackson CR, Barrett JB, Cleary T, Davidson M, Davis J, Mukherjee S, Fleming LE, Solo-Gabriele HM. 2013. Human-associated methicillin-resistant Staphylococcus aureus from a subtropical recreational marine beach. Microb. Ecol. 65:1039–1051. 10.1007/s00248-013-0216-1 [DOI] [PubMed] [Google Scholar]
  • 75.Harris SR, Feil EJ, Holden MT, Quail MA, Nickerson EK, Chantratita N, Gardete S, Tavares A, Day N, Lindsay JA, Edgeworth JD, de Lencastre H, Parkhill J, Peacock SJ, Bentley SD. 2010. Evolution of MRSA during hospital transmission and intercontinental spread. Science 327:469–474. 10.1126/science.1182395 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Murchan S, Aucken HM, O'Neill GL, Ganner M, Cookson BD. 2004. Emergence, spread, and characterization of phage variants of epidemic methicillin-resistant Staphylococcus aureus 16 in England and Wales. J. Clin. Microbiol. 42:5154–5160. 10.1128/JCM.42.11.5154-5160.2004 [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplemental material

Articles from Antimicrobial Agents and Chemotherapy are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES