Commentary: Comparative Genomic Analysis of Ten Clinical Streptococcus pneumoniae Collected From a Malaysian Hospital Reveal 31 New Unique Drug-Resistant SNPs Using Whole Genome Sequencing

Hassan Mahmood Jindal1, Shamala Devi Sekaran2*

1Department of Medical Microbiology, University of Malaya, Kuala Lumpur, 50603, Malaysia

2Department of Microbiology, Faculty of Medicine, MAHSA University, 42610 Jenjarom, Selangor, Malaysia


Despite the effort and decades of research, S. pneumoniae remains a primary cause of infectious morbidity and mortality worldwide. Although Antibiotics are lifesaving medications that offer tremendous benefits to patients with infectious diseases. Yet, several reports have revealed that the overuse and misuse of these agents had led to antibiotic resistance. Our study utilized whole genome sequencing (WGS) to reveal the pattern of antibiotic-resistance among ten pneumococcal isolates with various degree of susceptibility to antibacterial drugs. The main purpose of our study was to explore genetic variations related to drug-resistance in those ten strains. The results indicated that pneumococcal strains with resistant profile were associated with greater number of SNPs compared to susceptible ones. Out of all the SNPs identified, 31 were unique and had not been reported before. Our data propose that these SNPs could possess an important role in modifying the degree of sensitivity to different antibacterial drugs. In this article we comment on the methodology and results of our study which previously published in Journal of Biomedical Science.


Streptococcus pneumoniae or the pneumococcus is a major cause of community-acquired pneumonia and meningitis, as well as bloodstream, ear, and sinus infections1–4. Globally, it is estimated that this bacterial pathogen colonizes as many as 40–60% of young children. While colonization most often results in asymptomatic carriage, S. pneumoniae is still responsible for a substantial burden of disease5. In 2000, pneumococcus caused 14.5 million episodes of severe pneumococcal infections resulting in 826 000 deaths in children beneath the age of five6. Apart from pneumococcal deaths in HIV-positive children, death caused by pneumococcus accounts for approximately 11% of under-five mortality5. Decades of overuse of antibiotics in medical and agricultural applications as well as inappropriate prescribing of these drugs were the primary driver of antibiotic resistance crisis7–9. Drug-resistant S. pneumoniae (DRSP) has become an important clinical and public health problem during the past 20 years10. Pneumococcal resistance to different antibiotics led to 32,398 extra outpatient visits and 19,336 additional hospitalizations, accounting for $91 million (4%) in direct medical costs and $233 million (5%) in total costs, including work and productivity losses11. In order to overcome this issue and for better understanding on how pneumococci develop resistance to different types of antibiotics we utilized Whole Genome Sequencing (WGS) in our study12. WGS allow researchers to study the mode of action of antibiotics and the mechanisms involved in bacterial resistance13,14. Also, WGS can be applied by scientists to investigate the molecular basis and rate of evolution of antibiotic resistance in real-time under treatment regimens of single drugs or drugs combinations15. In our study, we used Whole Genome Sequencing to reveal the patterns of resistance of 10 pneumococcal isolates with a range of susceptibility and resistance to four different antibiotics: penicillin, cefotaxime, erythromycin, and tetracycline. The aim of our study was to investigate the genetic variation among pneumococcal isolates with different susceptibility profiles to four antibiotics in order to identify SNPs associated with virulent genes that could be a target for drug development.

The aim of our study is to identifiy Single Nucleotide Polymorphisms (SNPs) that are associated with antibiotic-resistance. Association of a SNP with drug resistance implicates genes that either reside near the genomic location of the SNP, or are regulated by a genetic factor located there. Ten clinical isolates of S. pneumniae were collected previously from University of Malaya Medical Centre (UMMC) (Table 1). The genomes of S. pneumniae clinical isolates were extracted using DNeasy Blood & Tissue Kit (Qiagen), the quantity and purity of the DNA was measured using qubit (Table 2). DNA fragmentation was done using Covaris S2. The fragmented DNA were ends repaired, added with dA base and ligated with Illumina indexed adapters. Standard concentration was used as the quantification becomes less reproducible, the sequencing library becomes less stable and subsequently Lower sequencing yield is the likely outcome. Size selections of the samples were performed using Invitrogen 2% agarose E-gels. The selected DNA fragments with adapters molecules on both ends underwent 10 cycles of PCR for amplification of prepared material. The samples were then diluted to 10Nm using hybridization buffer and pooled in to one pool. The libraries were loaded onto 1 lane of Illumina HiSeq 2000 flow cell v3 for sequencing.

Table 1: Bacterial strains and sources used for the genomic comparison of S. pneumniae strains.
Isolate Isolation date Sex Source Serotypea
SPS1 15/9/2010 NA Nasopharyngeal swab NT
SPS2 21/5/2011 Female Nasopharyngeal swab 1
SPS3 21/5/2011 Male Nasopharyngeal swab 19F
SPS4 20/2/2012 Female Nasopharyngeal swab 14
SPS5 16/3/2012 Female Swab from eye 23F
SPS6 18/5/2012 Male Nasopharyngeal swab 15B/C
SPS7 9/5/2011 Male Blood 1
SPS8 8/3/2011 Female Nasopharyngeal swab 14
SPS9 26/4/2011 Male Blood 18
SPS10 10/5/2011 Male Blood 8

a all serotypes were identified using multiplex PCR as described before (Pai et al., 2006).

Abbreviations: NA, not available; NT, non-typeable.

Table 2: Quantity of Samples using Qubit
Sample ID ng/ul Volume (ul) Total ng
SPS1 80.3 80 6424
SPS2 237 80 18960
SPS3 105 80 8400
SPS4 78.3 80 6264
SPS5 172 80 13760
SPS6 127 80 10160
SPS7 70.6 80 5648
SPS8 70.8 80 6372
SPS9 64.7 90 5823
SPS10 31.6 90 2844

In order to exclude low quality reads, PRINSEQ version 0.20.3 was used and the following types of reads were removed:

1. Reads having ‘N’ in more than 10% of the total bases of that read

2. Reads with Phred quality score less than 20.

3. Reads shorter than 50 bp.

In order to evaluate the core genome average identities and completeness, the sequenced reads were assembled and mapped against S. pneumniae TIGR4. SPAdes assembler was used in our study to assemble the genomic DNA extracted from the bacterial samples. This software is initially designed to assemble small genomes from MDA single-cell and standard bacterial data sets. Assembly of single cell data is challenging due to non-uniform read coverage, difference in insert length, high levels of sequencing errors and chimeric reads. Thus, SPAdes addresses these issues by performing assembly in four stages:

1. SPAdes proposes a new approach to assembly graph construction that uses the multisized de Bruijn graph, implementation of new bulge/tip removal algorithms, detection and removing of chimeric reads, aggregation of biread information into distance histograms, and allowing of backtrack the performed graph operations.

2. k-bimer adjustment, SPAdes derives accurate distance estimates between k-mers in the genome using joint analysis of distance histograms and paths in the assembly graph.

3. Paired assembly graph construction. By using k-bimer adjustment approach, SPAdes first extracts k-bimers from bireads, resulting in k-bimers with inexact distance estimates. The second step is transforming this set of k-bimers into a set of adjusted k-bimers with exact or almost exact distance estimates.

4. Contig construction.

To build a phylogenetic tree based on the identified SNPs, kSNP3 program was used. The reason for using this software over other available ones is that kSNP detects SNPs and builds phylogenies for large numbers of finished and draft sequences. Unlike other methods such as Parsnp which aligns the core genome and requires finished or assembled genomes, kSNP can use raw reads and is able to analyze hundreds of bacterial or viral genomes in only a few hours. In addition, kSNP can build Maximum Likelihood, Neighbor Joining, and parsimony phylogenetic trees based on all SNPs, only core SNPs, and SNPs present in at least a user-specified fraction of genomes. Realphy is another method to build a phylogenetic tree. This method maps raw reads to several reference genomes, therefore increasing the probability of using all of the information in the raw-read genomes for analysis. However, this method relays on accurate mapping of raw reads to the reference genomes, and if some taxa are diverged by > 5–10% the distances to the reference genome are under estimated, leading to incorrect topologies. kSNP overcomes this issue By not relaying on reference genome and by the ability of using raw read files.

Ten pneumococcal isolates with different sensitivity to four antibiotics were used in this study (Table 3). By using WGS we were able to found that the majority of the non-synonymous SNPs associated with pneumococcal essential genes were present in antibiotic resistant strains12. Through our analysis we were able to identify 90 non-synonymous SNPs related to the essential genes of the resistant strains, and some of them have reappeared in more than one resistant isolate, while none of these SNPs have occurred in susceptible isolates (Table 4). In addition, we were able to identify 31 unique SNPs associated with penicillin binding proteins, pneumolysin, PspA, sensor histidine kinase (ciaH) and capsular polysaccharide biosynthesis protein CpsA (Table 5). Phylogenetic analysis is the most commonly used tool to predict biological relationships. We used the parsimony tree to estimate the phylogenetic relationships among the clinical strains of S. pneumonaie. Our results are in agreement with the MIC profile of the ten pneumococcal strains. The observations that pneumococcal isolates with similar MIC profile were gathered together in the phylogenetic tree propose that these strains possess shared mutations and were probably originated from the same clone. It is possible that these strains could have evolved and acquired mutations in a similar manner due to selection pressures. The high phylogenetic relatedness among the clinical pneumococcal isolates with similar MIC profile is related to the specific SNPs in the mutated genes. The presence of identical uncommon mutations, as well as certain genes in the grouped isolates in the phylogenetic tree, is indicative of a single cluster of strains circulating in the population.

Table 3: Antibiotic susceptibility profiles of S. pneumonaie isolates.
Isolate a MIC (µg/ml)b
PEN c CTX c ERY c TET c
SPS1 2 1 2 16
SPS2 2 1 >2 >16
SPS3 4 1 >2 >16
SPS4 0.06 ≤0.063 ≤0.016 >16
SPS5 1 0.125 0.031 4
SPS6 0.06 ≤0.063 ≤0.016 ≤0.125
SPS7 2 2 >2 >16
SPS8 0.5 >8 2 >16
SPS9 0.25 8 2 16
SPS10 2 2 >2 16

a Isolates SPS1, SPS2, and SPS3 are non-susceptible to all antibiotics. Isolate SPS4 is susceptible to penicillin, cefotaxime, and erythromycin, but resistant to tetracycline. SPS5 is susceptible to cefotaxime and erythromycin, but resistant to penicillin and tetracycline. SPS6 is susceptible to all four antibiotics. SPS7 and SPS10 are resistant to all four antibiotics, SPS8 and SPS9 were resistant to all antibiotics except penicillin.

b MIC, Minimum inhibitory concentration.

c PEN, Penicillin; CTX, Cefotaxime; ERY, Erythromycin; TET, Tetracycline.

Table 4: Conserved non-synonymous Single Nucleotide Polymorphisms (SNPs) associated with Penicillin Binding Proteins (PBPs) and other virulent genes found in resistant isolates.
Locus Name Putative Identification Reference
Position
TIGR4 SNP Pneumococcal isolate Amino Acid Change
SP_0346 cpsA; capsular polysaccharide biosynthesis protein 320234 C T SPS7, SPS8, SPS10 A53V
320204 C T SPS7, SPS8 A43V
320872 C T SPS10 P266S
321451 G A SPS10 V459M
320582 C T SPS10 A169V
320657 C T SPS9, SPS10 S194L
320560 A G SPS10 N162D
321410 T C SPS9 M445T
320314 G C SPS1, SPS9 V80L
321460 A G SPS2 I462V
321485 T C SPS7, SPS8 V470A
320101 C A SPS2 R9S
320710 A G SPS1 T212A
SP_0347 cpsB; capsular polysaccharide biosynthesis protein Cps4B 321798 A G SPS2, SPS10 E92G
321954 A G SPS10 E144G
322094 G A SPS10 D191N
322169 C T SPS10 L216F
322235 G A SPS10 V238I
321579 G T SPS9 R19I
321608 T G SPS1, SPS2, SPS7, SPS8 S29A
SP_0348 cpsC; capsular polysaccharide biosynthesis protein 322306 G T SPS2, SPS10 V15F
322313 G C SPS10 S17T
322321 A G SPS10 K20E
322342 A T SPS2, SPS10 I27L
322456 C T SPS10 P65S
322489 A T SPS10 T76S
322853 A T SPS1, SPS7, SPS8 H197L
322360 G A SPS2 G33S
322693 G A SPS7, SPS8 E144K
322549 G A SPS9 V96I
SP_0349 cpsD; capsular polysaccharide biosynthesis protein 323314 G A SPS9 V117I
323202 G A SPS9 M79I
323488 G A SPS1, SPS7, SPS8 V175I
323191 A C SPS1, SPS7, SPS8 N76H
323193 T A SPS1, SPS7, SPS8 N76K
323416 A G SPS1, SPS7, SPS8, SPS9 I151V
SP_1837 capsular polysaccharide biosynthesis protein 1746914 T C SPS1, SPS9, SPS10 K212R
1747016 A G SPS1, SPS2, SPS7, SPS8, SPS9, SPS10 I178T
1747484 A G SPS1, SPS2, SPS7, SPS8, SPS9, SPS10 V22A
1747494 T C SPS1, SPS2, SPS7, SPS8, SPS9, SPS10 T19A
SP_0117 pspA; pneumococcal surface protein A 118489 A G SPS1, SPS7, SPS8, SPS9 T23A
118490 C T SPS9 T23M
120628 A G SPS9, SPS10 K736E
120431 C A SPS7, SPS8 A670D
119178 A C SPS7, SPS8 K252N
119449 A C SPS7, SPS8 K343Q
119056 T G SPS1 Y212D
118496 A C SPS7, SPS8, SPS10 Q25P
SP_0799 ciaH; sensor histidine kinase ClaH 753163 C G SPS7, SPS8 H180D
SP_1923 pln; pneumolysin 1832851 G A SPS9 T154M
1832174 T C SPS2, SPS9, SPS10 N380D
1832641 T C SPS10 K224R
1832797 G A SPS10 T172I
1831975 G A SPS7, SPS8 P446L
1832906 G T SPS1, SPS7, SPS8 Q136K
SP_1937 lytA; autolysin 1840479 T G SPS2 L295I
1840604 A T SPS2 E253V
1840608 C T SPS2 N252D
1840473 A G SPS9 P297S
1840624 A C SPS1 D246E
SP_0369 penicillin-binding protein 1A 347449 C T SPS1 A522T
347857 C T SPS1 V386I
347479 C T SPS2 E512K
348706 T A SPS2 T103S
347473 C G SPS10 E514Q
SP_2099 penicillin-binding protein 1B 2006807 A G SPS10 V787A
2007578 T G SPS9 E530A
SP_2010 penicillin-binding protein 2A 1917863 T C SPS9, SPS10 E17G
1917045 T C SPS9, SPS10 T290A
1916273 C T SPS1, SPS9, SPS10 S547N
1916459 T G SPS9 A485E
1916166 C T SPS9 A583T
1917111 G T SPS2 Q268K
1916595 C T SPS2 D440N
1916819 A G SPS1 F365S
SP_1673 penicillin-binding protein 2B 1573249 C T SPS7, SPS8, SPS9, SPS10 G597E
1573212 C A SPS9 L609F
1574933 C T SPS3 V36I
1573493 C A SPS2 A516S
1574288 C A SPS2, SPS3 A251S
1574461 G A SPS2 A193V
SP_0336 penicillin-binding protein 2X 309007 C T SPS9, SPS10 L710F
308341 G A SPS9 D488N
307393 G A SPS2, SPS3 A172T
309113 C A SPS3 T745K
SP_0798 ciaR; DNA-binding response regulator 751980 G A SPS2 V7I
SP_0377 cbpC; choline-binding protein C 356412 G A SPS1 G156S
355972 A C SPS2 Q9P
355974 G A SPS2 V10I
356806 C T SPS2 S287L
356182 C A SPS7, SPS8 P79H
356044 A G SPS10 R33Q
Table 5: Unique non-synonymous Single Nucleotide Polymorphisms (SNPs) associated with Penicillin Binding Proteins (PBPs) and other virulent genes found in all ten pneumococcal isolates isolates.
Locus Name Putative Identification Reference
Position
TIGR4 SNP Pneumococcal isolate Amino Acid Change
SP_0346 cpsA; capsular polysaccharide biosynthesis protein 320234 C T SPS7, SPS8, SPS10 A53V
320204 C T SPS7, SPS8 A43V
320872 C T SPS10 P266S
321451 G A SPS10 V459M
320657 C T SPS9, SPS10 S194L
321460 A G SPS2 I462V
321485 T C SPS7, SPS8 V470A
320710 A G SPS1 T212A
322321 A G SPS10 K20E
322360 G A SPS2 G33S
323191 A C SPS1, SPS7, SPS8 N76H
SP_1837 capsular polysaccharide biosynthesis protein 1746914 T C SPS1, SPS9, SPS10 K212R
SP_0117 pspA; pneumococcal surface protein A 118490 C T SPS9 T23M
120431 C A SPS7, SPS8 A670D
118496 A C SPS7, SPS8, SPS10 Q25P
SP_0799 ciaH; sensor histidine kinase ClaH 753163 C G SPS7, SPS8 H180D
SP_1923 pln; pneumolysin 1832851 G A SPS9 T154M
SP_0369 penicillin-binding protein 1A 347479 C T SPS2 E512K
348706 T A SPS2 T103S
SP_2099 penicillin-binding protein 1B 2006807 A G SPS10 V787A
SP_2010 penicillin-binding protein 2A 1917863 T C SPS9, SPS10 E17G
1916459 T G SPS9 A485E
1916166 C T SPS9 A583T
SP_1673 penA; penicillin-binding protein 2B 1573212 C A SPS9 L609F
1574933 C T SPS3 V36I
1573493 C A SPS2 A516S
1574288 C A SPS2, SPS3 A251S
1574461 G A SPS2 A193V
SP_0336 penicillin-binding protein 2X 308341 G A SPS9 D488N
307393 G A SPS2, SPS3 A172T
309113 C A SPS3 T745K

In summary, we compared the genomic sequences of ten pneumococcal strains isolated from University of Malaya Medical Centre (UMMC) with different sensitivity to four different antibiotics: penicillin, cefotaxime, erythromycin, and tetracycline in order to identify the genetic variations within the sequences of these isolates and identifying SNPs that could play significant role in conferring resistance to those antibiotics. The high level of sequence conservation and the presence of the same mutations mainly those associated with genes involved in β-lactam resistance in both sensitive and resistant isolates makes it a difficult task to identify distinct mechanisms of resistance that differentiate strains with different drug-sensitivities, and that antibiotic resistance cannot be only linked to the presence of certain genes. Nevertheless, through our extensive analysis we were able to identify unique SNPs associated with virulent genes that could play a key role in resistance to various antibiotics. However, the small number of the clinical samples included in this study has limited our understanding to the role of these SNPs in conferring resistance toward different antibiotics. Moreover, all resistant genes have yet to be subjected to individual mutational analysis. This can be achieved by introducing the identified SNPs to the resistant genes by site-directed mutagenesis and further expression analysis to confirm the role of these SNPs in conferring antibiotic resistance.

All tables included in this commentary article are reproduced from the original research article published in Journal of Biomedical Science (2018) 25:15.

  1. Kim L, McGee L, Tomczyk S, et al. Biological and epidemiological features of antibiotic-resistant Streptococcus pneumoniae in pre- and post-conjugate vaccine eras: A United States perspective. Clin Microbiol Rev. 2016; 29(3): 525-552. doi:10.1128/CMR.00058-15.
  2. Jindal HM, Le CF, Yusof MYM, et al. Antimicrobial activity of novel synthetic peptides derived from indolicidin and ranalexin against Streptococcus pneumoniae. PLoS One. 2015; 10(6): e0128532. doi:10.1371/journal.pone.0128532.
  3. Jindal HM, Zandi K, Ong KC, et al. Mechanisms of action and in vivo antibacterial efficacy assessment of five novel hybrid peptides derived from Indolicidin and Ranalexin against Streptococcus pneumoniae. PeerJ. 2017; 5(10): e3887. doi:10.7717/peerj.3887.
  4. van der Poll T, Opal SM. Pathogenesis, treatment, and prevention of pneumococcal pneumonia. Lancet. 2009; 374(9700): 1543-1556. doi:10.1016/S0140-6736(09)61114-4.
  5. Mitchell PK, Lipsitch M, Hanage WP. Carriage burden , multiple colonization and antibiotic pressure promote emergence of resistant vaccine escape pneumococci. 2015.
  6. O’Brien KL, Wolfson LJ, Watt JP, et al. Burden of disease caused by Streptococcus pneumoniae in children younger than 5 years: global estimates. Lancet. 2009; 374(9693): 893-902. doi:10.1016/S0140-6736(09)61204-6.
  7. Ventola CL. The antibiotic resistance crisis: part 1: causes and threats. P T A peer-reviewed J Formul Manag. 2015; 40(4): 277-83. doi:Article.
  8. Cornick JE, Bentley SD. Streptococcus pneumoniae: the evolution of antimicrobial resistance to beta-lactams, fluoroquinolones and macrolides. Microbes Infect. 2012; 14(7-8): 573-83. doi:10.1016/j.micinf.2012.01.012.
  9. Shea KM. Antibiotic resistance: what is the impact of agricultural uses of antibiotics on children’s health. Pediatrics. 2003; 112(1): 253-258. doi:10.1542/peds.112.1.S1.253.
  10. Nuermberger EL, Bishai WR. Antibiotic Resistance in Streptococcus pneumoniae: What Does the Future Hold. Clin Infect Dis. 2004; 38(Supplement 4): S363-S371. doi:10.1086/382696.
  11. Reynolds C a, Finkelstein J a, Ray GT, et al. Attributable healthcare utilization and cost of pneumonia due to drug-resistant streptococcus pneumonia: a cost analysis. Antimicrob Resist Infect Control. 2014; 3(1): 16. doi:10.1186/2047-2994-3-16.
  12. Jindal HM, Ramanathan B, Le CF, et al. Comparative genomic analysis of ten clinical Streptococcus pneumoniae collected from a Malaysian hospital reveal 31 new unique drug-resistant SNPs using whole genome sequencing. J Biomed Sci. 2018; 25(1). doi:10.1186/s12929-018-0414-8.
  13. Köser CU, Ellington MJ, Peacock SJ. Whole-genome sequencing to control antimicrobial resistance. Trends Genet. 2014; 30(9): 401-407. doi:10.1016/j.tig.2014.07.003.
  14. Cui L, Neoh H, Shoji M, et al. Contribution of vraSR and graSR point mutations to vancomycin resistance in vancomycin-intermediate Staphylococcus aureus. Antimicrob Agents Chemother. 2009; 53(3): 1231-4. doi:10.1128/AAC.01173-08.
  15. Palmer AC, Kishony R. Understanding, predicting and manipulating the genotypic evolution of antibiotic resistance. Nat Rev Genet. 2013; 14(6): 243-248. doi:10.1016/j.biotechadv.2011.08.021.Secreted.
 

Article Info

Article Notes

  • Published on: August 08, 2018

Keywords

  • Antibiotic-resistance

  • Whole Genome Sequencing
  • Single Nucleotide Polymorphism

*Correspondence:

Prof. Shamala Devi Sekaran
Faculty of Medicine, MAHSA University, Saujana Putra Campus Jalan SP 2, Bandar Saujana Putra, 42610 Jenjarum, Selangor, Malaysia
Telephone No: +603 510 222 98; Fax No: +603 7931 7118
Email: shamalamy@yahoo.com.