• No results found

University of Groningen Quantifying the transcriptome of a human pathogen Aprianto, Rieza

N/A
N/A
Protected

Academic year: 2021

Share "University of Groningen Quantifying the transcriptome of a human pathogen Aprianto, Rieza"

Copied!
15
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

University of Groningen

Quantifying the transcriptome of a human pathogen

Aprianto, Rieza

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from

it. Please check the document version below.

Document Version

Publisher's PDF, also known as Version of record

Publication date:

2018

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Aprianto, R. (2018). Quantifying the transcriptome of a human pathogen: Exploring transcriptional

adaptation of Streptococcus pneumoniae under infection-relevant conditions. Rijksuniversiteit Groningen.

Copyright

Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policy

If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum.

(2)

6

Discussion

(3)
(4)

6

RN

A: mor

e than a middle man

RNA: more than a middle man

The central dogma states that a gene, made of DNA, transcribes informa-tion to a transcript made of RNA, which in turn is translated into a protein, the main cellular work-horse1. The dogma insinuates that the transcript

is a simple baton between the repository of information and the effector. Actually, RNA can directly regenerate a copy of itself, also known as RNA replication. Moreover, RNA can transcribe DNA, reverting the dogmatic information flow. On the other hand, novel developments in sequencing technology and recent studies have revealed an expanded RNA function in the regulation of gene expression, including modifying the abundance of transcripts and efficiency of translation. Finally, the genome-wide quantification of transcripts has been made easy by the advent of next generation sequencing.

The opportunistic human pathogen, Streptococcus pneumoniae, em-ploys manifold strategies to orchestrate its gene expression. An import-ant class, the stand-alone regulator, is mediated by a protein binding directly to DNA to regulate downstream expression. For example, CcpA, the carbon catabolite protein, has a sensing domain and a DNA-binding domain, moonlighting two roles in one molecules2. Another type of a

pro-tein-based regulator, the two component systems (TCSs), separate these functions by a membrane-bound sensor and a cytoplasmic DNA-binding regulator. Specifically, TCSs architecture makes it suitable to detect envi-ronmental signals3.    

On the other hand, pneumococcal RNA-based regulation has not been fully appreciated with only intermittent studies available4. This is wholly

surprising, since RNA-mediated gene regulation is highly suitable for

S. pneumoniae facing a dynamic environment during colonization,

trans-mission and infection. First, RNA as gene regulator is more economical to the cell than expressing a protein to control expression, simply because transcribing RNA does not require an investment into amino acids5. For

example, high-cost enzymes are mostly controlled at the transcriptional level in other bacteria6, emphasizing resource-prudent decision-making

in order to conserve cellular resources, such as amino acids and tRNAs. Secondly, RNA molecules are generally less stable than a protein7, highly

(5)

166

6

Discus

sion

suitable for mediating transient regulation. Recent transcriptomics stud-ies further support this hypothesis by highlighting the ubiquity of pneu-mococcal non-coding small RNAs and thus, the potential of RNA-based

regulations8–10. Pneumococcal gene regulation may also benefit by

com-bining RNA- and protein-based regulators, for example, five sRNAs under the control of CiaRH have been reported to have wide-ranging effects on

pneumococcal phenotypes11,12. Lastly, pneumococcal riboswitches have

not been formally described, aside from orthology-based reports4.

These central roles of transcripts in the cellular function of S.

pneumo-niae including in the development of pathogenesis served as the context

of our study into the pneumococcal genomic repertoire. In Chapter 2, we

employed cutting-edge sequencing technology to thoroughly define the pneumococcal genome-transcriptome and to revise the genome-wide annotation, including non-coding RNAs. Furthermore, we reported abundant transcription from the antisense strand (asRNAs), overlapping coding sequences and untranslated regions. The abundance of antisense RNAs can directly regulate gene expression in a cis-acting manner. In ad-dition, we observed significant transcriptional starting sites (TSS, +1) in the middle of the annotated coding sequence or putative operon. The presence of internal transcriptional start sites (iTSSs) may suggest an al-ternative regulation of genes within operons. Even more, antisense and internal transcription might be activated in response to certain environ-mental signals as an adaptive measure to expand phenotypic plasticity.

Furthermore, in Chapter 2, we have re-categorized four small genomic

features as uridine-responsive riboswitches. During uridine abundance, the riboswitch destabilizes an antiterminator hairpin, causing early termi-nation and switching off the downstream expression13. Moreover, a recent

study showed an early termination to be independent of PyrR, a repres-sor previously thought to be essential in facilitating riboswitch function14.

Riboswitch-based gene regulation has been proposed to be more bene-ficial in Gram positive than Gram negative bacteria, due to the generally more compact genome and longer operons in Gram positive bacteria5. For

instance, Bacillus subtilis, a Gram positive model organism, puts 2% of its genes under the control of riboswitches5. In addition, regulation by

(6)

6

Quan

tifying inf

ection

of gene expression, at both the transcriptional and translational levels15.

Lastly, RNA fragments originating from early termination events may act as trans-acting sRNA16, further moonlighting its regulatory function. In

all, our genome and curated annotation will provide the chassis for future work on pneumococcal genomics. At the same time, our work contrib-uted to the perspective of RNA-mediated gene regulation in S.

pneumo-niae. Chapter 2 re-emphasizes the role of RNA-based gene regulation in S. pneumoniae by highlighting the genomic potential of RNA features

in-cluding riboswitches, UTRs and sRNAs in fine tuning selected transcrip-tion and translatranscrip-tion.

Quantifying infection: genome-wide transcript abundance

in models

The glaring lack of pathogenicity islands suggests that pneumococcal virulence and further pathogenesis are more subtly regulated17–19, for

ex-ample via mutation in untranslated regions (UTRs). In addition, recent advances in sequencing technology have allowed precise quantification of transcript abundance20. In Chapter 3, we combined the quantification

of total RNA with an array of in vitro infection-relevant models, persuad-ing the bacteria to express a compendium of expression in response to a wide range of microenvironments. The approach has been applied to other relevant bacteria21–24, usually in conjunction with the elucidation of

start sites. In this study, we observed that transcriptional levels of genes involved in core cellular functions, such as carbon metabolism, transcrip-tion and translatranscrip-tion, are conserved across varying conditranscrip-tions.

In this study, we did not deplete bacterial rRNA thereby avoiding the bias introduced by depletion25. Here, we have observed that rRNAs occupy

around 95% of transcripts while tRNAs take up less than 1% of total tran-scripts. Subsequently, transcript abundance was transformed into a  co- expression matrix, a form of gene network. We combined correlation val-ues of neighboring genes, start sites and terminator sites (from Chapter 2)

to map genome-wide putative operons. The co-expression matrix may serve as a base to build gene regulatory networks by inferring regula-tion among genomic features. By exploiting this method, we revealed

(7)

168

6

Discus

sion

gene had not been observed as part of the competence regulon in pre-vious array-based studies26,27, exemplifying the power of next- generation

sequencing in transcriptomics quantification.  

In Chapter 4, we used RNA-sequencing to determine a suitable

re-porter gene to tag with a fluorescent protein. We selected HlpA, a highly transcribed protein and the only histone-like protein in Streptococcus28.

Furthermore, we followed the fluorescent pneumococcal reporter strain in several infection-relevant models, including during co-incubation with live confluent human lung epithelial cells. The exposure of S. pneumoniae to confluent epithelial layers has been shown to reflect a biotic surface as present in vivo, and it stimulates different phenotypic responses com-pared to growth in liquid broth or growth on a plastic abiotic surface29.

Moreover, we included an unencapsulated strain in the study. The unen-capsulated derivative showed better adherence to epithelial cells than the ancestral encapsulated strain. In addition, the pneumococcal polysaccha-ride capsule is an essential virulence factor that determines colonization and infection success; it also serves as a vaccine target30.

We then exploited the co-incubation model including the unencap-sulated strain to simultaneously sequenced the host and pathogen tran-scripts during the progression of early infection (Chapter 5). Since the

unencapsulated strain better adheres to epithelial cells, these bacteria have more intimate contact with the monolayer than its encapsulated, mostly planktonic, strain. By comparing transcriptional responses between unen-capsulated and wild type models, we elucidated adherence-specific tran-scriptional responses in both epithelial and pneumococcal transcriptomes. Because adherence is the first step towards pathogenesis, adherence- specific regulated genes might provide interesting novel drug and/or vac-cination targets that have not yet been identified by conventional screens aimed at targeting essential and/or surface exposed proteins. Nevertheless, future work to precisely identify useful targets is required.

In Chapter 5, we have developed a rapid RNA isolation technique to

ter-minate the protein-based RNA production and digestion allowing precise kinetics observations. Furthermore, in cooperation with EMBL Genomics Core Facility, we developed a one-step dual rRNA depletion to remove both human and bacterial-derived rRNAs. We further exploited the bright

(8)

6

Quan

tifying inf

ection

fluorescent strain (from Chapter 4) to visualize epithelial- adherent

bacte-ria during the progression of early infection. In addition, we showed that the activity of promoter-of-interest can be measured by fluorescence tag-ging and live microscopy. Also, we have shown that the adherent bacteria modulate the innate immune response of epithelial cells to their advan-tage while importing host-associated sugars as a carbohydrate source to fuel bacterial expansion.

Remarkably, we observed that all pneumococcal genes are expressed at one point or another during the infection process (Chapter 5).

We no-ticed the same observation in Chapter 3 when the bacteria were exposed

to conditions relevant to its lifestyle. While this might partly stem from the high sequencing depth undertaken in these studies, other genome- wide bacterial gene expression studies confirm the abundant bacterial tran-scription21,22. We speculate that dynamic conditions, including interspecies

interaction, necessitates massive pneumococcal transcriptional adapta-tion. Additionally, we have also shown that the whole of the genome can be transcribed due to ubiquitous transcriptional start sites (Chapter 2). To

summarize, Chapters 3 and 5 highlights the suitability of transcriptomics

(i.e. genome-wide quantification of transcriptional responses) in elucidat-ing the highly complicated interaction between the bacteria and its im-mediate microenvironment, both abiotic and biotic components, during the progression of colonization and infection. High throughput RNA-seq, combined with suitable harvesting methods offers a snapshot into the continuous interaction between S. pneumoniae and the components of microenvironment.

Furthermore, in all chapters, we have strived to disseminate the results of our research as widely as possible. First and foremost, we have been using open access platforms to publish our results. It should be noted that transcriptome databases only exist for few model bacteria such as Bacillus  subtilis22,31, Staphylococcus aureus32, Escherichia coli33,34 and

Salmonella enterica Serovar Typhimurium35. Despite their paucity, these

resources have been proven to be invaluable resources for the research com-munity. Thereby, we have built easy-to-access online browsers for our data-sets (Chapters 2, 3 and 5) where users can simply choose a gene of interest

(9)

170

6

Discus

sion

The future: time to complicate matters

In the near future, further maturation of sequencing technologies prom-ises higher quality and more opportunities. Currently, indispensable sample preparations such as amplification and reverse transcription in-troduce bias to the data25; direct sequencing may remove these biases by

bypassing the sample preparations altogether36. Furthermore, current

technologies average individual responses into an aggregate population response. On the other hand, individual transcriptional changes in bacte-ria represents a meaningful diversity of responses, which, in turn, drives varying host responses37. Thus, a combination of direct sequencing and

single cell transcriptomics will open novel and powerful approaches in elucidating pneumococcal pathogenesis and other infectious agents. In addition, while genome-wide quantification of transcripts gives global impression of cellular protein abundance, the correlation between RNA and protein abundance is still limited (~0.7 to 0.8, in E. coli)38. Thus,

tran-scriptomics approaches benefit from complementary approaches, includ-ing proteomics and metabolomics.

Furthermore, the nasopharyngeal passage is a bustling ecosystem with numerous pneumococcal strains and onther bacteria39,40. Multiple strains

of S. pneumoniae can be present in the same individual host while occu-pying and competing for niche and resources. Furthermore, interstrain

competition has been enhanced by the pressure from vaccinations41 by

the replacement of target strains with non-vaccine serotypes. In addi-tion, while a study of the genomes of 44 clinical and model strains indi-cates that up to  74% of the genome is conserved42, strain-specific

tran-scriptional response to the colonization and infection sites may be more nuanced. The soon-available data on pneumococcal core genome and pan-genome from the Global Pneumococcal Sequencing project, where more than 20,000 pneumococcal genomes have already been sequenced (http://www.pneumogen.net/gps/) may serve as stepping stone to eluci-date the strain-specific responses.

Moreover, colonization and infection sites retain multispecies residents. S. pneumoniae competes for location and resources with

(10)

6

The fu

tur

e

influenzae, another common resident of the nasopharyngeal niche,

recruits host immune cells to hinder pneumococcal outgrowth44. Also,

dispersion biofilm models of either S. pneumoniae or S. aureus behave significantly different than a dual-species model, with co-colonization models causing early dispersion of S. pneumoniae and ending with

pneumococcal pneumonia45. Furthermore, in a murine co-infection

model of S. pneumoniae and influenza virus, pneumococci were found to utilize host- derived sugars to boost their own metabolism46. The

sug-ars were available as a  result of the activity of viral enzymes. In these examples, the interaction is not necessarily direct but meaningfully changes the nature of pneumococcal virulence and pathogenesis in the host. In addition, the interaction within a polymicrobial environment has been reported to elicit a different set of essential genes in bacteria

altogether47. These reports highlight the complex system of

microbi-ota niches, including the respiratory tract, and interspecies interaction in infection. In order to explore host- pneumococcal interaction in the whole organism for real in vivo dual RNA-seq, we are investigating tran-scriptional responses in meningeal infection with zebrafish embryo of both host and pathogen responses. There, we are combining the pre-viously described infection model48 and a strain with inactivated

viru-lence factor, pneumolysin. Finally, we postulate that the combined ap-proach of multispecies infection model and avant-garde development of sequencing technology will uncover more clinically-relevant patho-genic interactions. The knowledge and insights ultimately may lead to novel vaccination strategies or new targets that can be used to prevent the expression of pneumococcal virulence factor and stop its pathogen-esis inside the human host.

References

1. Crick, F. Central dogma of molecular biology. Nature 227, 561–563 (1970).

2. Kietzman, C. C. & Rosch, J. W. Regulatory Strategies of the Pneumococcus. in

Strepto-coccus pneumoniae: Molecular Mechanism of Host-Pathogen Interactions (eds. Brown,

J., Hammerschmidt, S. & Orihuela, C.) 109–128 (Academic Press, 2015). doi:10.1016/ B978-0-12-410530-0.00006-5

(11)

172

6

Discus

sion

3. Gomez-Mejia, A., Gamez, G. & Hammerschmidt, S. Streptococcus pneumoniae two-compo-nent regulatory systems: The interplay of the pneumococcus with its environment. Int. J.

Med. Microbiol. (2017). doi:10.1016/j.ijmm.2017.11.012

4. Wilton, J., Acebo, P., Herranz, C., Gómez, A. & Amblar, M. Small regulatory RNAs in

Strepto-coccus pneumoniae: discovery and biological functions. Front. Genet. 6, (2015).

5. Waters, L. S. & Storz, G. Regulatory RNAs in bacteria. Cell 136, 615–628 (2009).

6. Wessely, F. et al. Optimal regulatory strategies for metabolic pathways in Escherichia coli depending on protein costs. Mol. Syst. Biol. 7, 515 (2011).

7. Selinger, D. W., Saxena, R. M., Cheung, K. J., Church, G. M. & Rosenow, C. Global RNA half-life analysis in Escherichia coli reveals positional patterns of transcript degradation. Genome

Res. 13, 216–223 (2003).

8. Kumar, R. et al. Identification of novel non-coding small RNAs from Streptococcus

pneumo-niae TIGR4 using high-resolution genome tiling arrays. BMC Genomics 11, 350 (2010).

9. Mann, B. et al. Control of virulence by small RNAs in Streptococcus pneumoniae. PLoS

Pat-hog. 8, e1002788 (2012).

10. Tsui, H.-C. T. et al. Identification and characterization of noncoding small RNAs in

Strepto-coccus pneumoniae serotype 2 strain D39. J. Bacteriol. 192, 264–279 (2010).

11. Halfmann, A., Kovács, M., Hakenbeck, R. & Brückner, R. Identification of the genes directly controlled by the response regulator CiaR in Streptococcus pneumoniae: five out of 15 pro-moters drive expression of small non-coding RNAs. Mol. Microbiol. 66, 110–126 (2007).

12. Marx, P., Nuhn, M., Kovács, M., Hakenbeck, R. & Brückner, R. Identification of genes for small non-coding RNAs that belong to the regulon of the two-component regulatory system CiaRH in Streptococcus. BMC Genomics 11, 661 (2010).

13. Martinussen, J., Schallert, J., Andersen, B. & Hammer, K. The pyrimidine operon pyr-RPB-carA from Lactococcus lactis. J. Bacteriol. 183, 2785–2794 (2001).

14. Tanaka, Y., Teramoto, H. & Inui, M. Regulation of the expression of de novo pyrimidine bio-synthesis genes in Corynebacterium glutamicum. J. Bacteriol. 197, 3307–3316 (2015).

15. Serganov, A. & Nudler, E. A decade of riboswitches. Cell 152, 17–24 (2013).

16. Loh, E. et al. A trans-acting riboswitch controls expression of the virulence regulator PrfA in

Listeria monocytogenes. Cell 139, 770–779 (2009).

17. Blomberg, C. et al. Pattern of accessory regions and invasive disease potential in

Streptococ-cus pneumoniae. J. Infect. Dis. 199, 1032–1042 (2009).

18. Obert, C. et al. Identification of a candidate Streptococcus pneumoniae core genome and regions of diversity correlated with invasive pneumococcal disease. Infect. Immun. 74, 4766–

4777 (2006).

19. Silva, N. A. et al. Genomic diversity between strains of the same serotype and multilocus sequence type among pneumococcal clinical isolates. Infect. Immun. 74, 3513–3518 (2006).

20. Creecy, J. P. & Conway, T. Quantitative bacterial transcriptomics with RNA-seq. Curr. Opin.

Microbiol. 23, 133–140 (2015).

21. Kroger, C. et al. An infection-relevant transcriptomic compendium for Salmonella enterica Serovar Typhimurium. Cell Host Microbe 14, 683–695 (2013).

22. Nicolas, P. et al. Condition-dependent transcriptome reveals high-level regulatory architec-ture in Bacillus subtilis. Science 335, 1103–1106 (2012).

(12)

6

R ef er enc es

23. Sharma, C. M. et al. The primary transcriptome of the major human pathogen Helicobacter

pylori. Nature 464, 250–255 (2010).

24. Toledo-Arana, A. et al. The Listeria transcriptional landscape from saprophytism to viru-lence. Nature 459, 950–956 (2009).

25. Lahens, N. F. et al. IVT-seq reveals extreme bias in RNA sequencing. Genome Biol. 15, R86 (2014).

26. Dagkessamanskaia, A. et al. Interconnection of competence, stress and CiaR regulons in

Streptococcus pneumoniae: competence triggers stationary phase autolysis of ciaR mutant

cells. Mol. Microbiol. 51, 1071–1086 (2004).

27. Peterson, S. N. et al. Identification of competence pheromone responsive genes in

Strepto-coccus pneumoniae by use of DNA microarrays. Mol. Microbiol. 51, 1051–1070 (2004).

28. Liu, D. et al. The essentiality and involvement of Streptococcus intermedius histone-like DNA-binding protein in bacterial viability and normal growth. Mol. Microbiol. 68, 1268–1282

(2008).

29. Marks, L. R., Parameswaran, G. I. & Hakansson, A. P. Pneumococcal interactions with epithe-lial cells are crucial for optimal biofilm formation and colonization in vitro and in vivo. Infect.

Immun. 80, 2744–2760 (2012).

30. Geno, K. A. et al. Pneumococcal capsules and their types: past, present, and future. Clin.

Mi-crobiol. Rev. 28, 871–899 (2015).

31. Michna, R. H., Zhu, B., Mäder, U. & Stülke, J. SubtiWiki 2.0 - an integrated database for the model organism Bacillus subtilis. Nucleic Acids Res. 44, D654-662 (2016).

32. Gopal, T., Nagarajan, V. & Elasri, M. O. SATRAT: Staphylococcus aureus transcript regulatory network analysis tool. PeerJ 3, (2015).

33. Chang, X. et al. EcoBrowser: a web-based tool for visualizing transcriptome data of

Esche-richia coli. BMC Res. Notes 4, 405 (2011).

34. Ishii, N. et al. Multiple high-throughput analyses monitor the response of E. coli to perturba-tions. Science 316, 593–597 (2007).

35. Kroger, C. et al. The transcriptional landscape and small RNAs of Salmonella enterica sero-var Typhimurium. Proc. Natl. Acad. Sci. U. S. A. 109, E1277-1286 (2012).

36. Garalde, D. R. et al. Highly parallel direct RNA sequencing on an array of nanopores. Nat.

Methods (2018). doi:10.1038/nmeth.4577

37. Avraham, R. et al. Pathogen cell-to-cell variability drives heterogeneity in host immune re-sponses. Cell 162, 1309–1321 (2015).

38. Liu, Y., Beyer, A. & Aebersold, R. On the dependency of cellular protein levels on mRNA abundance. Cell 165, 535–550 (2016).

39. Bosch, A. A. T. M. et al. Development of upper respiratory tract microbiota in infancy is af-fected by mode of delivery. EBioMedicine 9, 336–345 (2016).

40. Bosch, A. A. T. M. et al. Nasopharyngeal carriage of Streptococcus pneumoniae and other bacteria in the 7th year after implementation of the pneumococcal conjugate vaccine in the Netherlands. Vaccine 34, 531–539 (2016).

41. Henriques-Normark, B. & Tuomanen, E. I. The pneumococcus: epidemiology, microbiology, and pathogenesis. Cold Spring Harb. Perspect. Med. 3, (2013).

42. Donati, C. et al. Structure and dynamics of the pan-genome of Streptococcus pneumoniae and closely related species. Genome Biol. 11, R107 (2010).

(13)

174

6

Discus

sion

43. Man, W. H., de Steenhuijsen Piters, W. A. A. & Bogaert, D. The microbiota of the respiratory tract: gatekeeper to respiratory health. Nat. Rev. Microbiol. 15, 259–270 (2017).

44. Lysenko, E. S., Lijek, R. S., Brown, S. P. & Weiser, J. N. Within-host competition drives se-lection for the capsule virulence determinant of Streptococcus pneumoniae. Curr. Biol. 20,

1222–1226 (2010).

45. Reddinger, R. M., Luke-Marshall, N. R., Sauberan, S. L., Hakansson, A. P. & Campagnari, A. A. Streptococcus pneumoniae modulates Staphylococcus aureus biofilm dispersion and the transition from colonization to invasive disease. mBio 9, e02089-17 (2018).

46. Siegel, S. J., Roche, A. M. & Weiser, J. N. Influenza promotes pneumococcal growth during coinfection by providing host sialylated substrates as a nutrient source. Cell Host Microbe 16,

55–67 (2014).

47. Ibberson, C. B. et al. Co-infecting microorganisms dramatically alter pathogen gene essenti-ality during polymicrobial infection. Nat. Microbiol. 2, 17079 (2017).

48. Jim, K. K. et al. Infection of zebrafish embryos with live fluorescent Streptococcus

(14)
(15)

Referenties

GERELATEERDE DOCUMENTEN

Annotated elements were divided over six annotation tracks: (i) genes (includes pseudogenes, shown in grey), (ii) putative operons, (iii) TSSs and terminators, (iv) predicted

Exploring transcriptional adaptation of Streptococcus pneumoniae under infection-relevant

compendium of pneumococcal transcriptome by exposing the strain to conditions relevant to the bacterial lifestyle of colonization and infec-. tion

SMRT (PacBio) Illumina single-end stranded Illumina paired-end stranded Mapping reads on genome β (bowtie2) de novo assembly (HGAP3) DNA methylation analysis Enriched 5’ base

tion-relevant conditions, a subset of genes was constantly highly ex- pressed while there is no gene that is always lowly expressed - highlighting the saturated and dynamic nature

By dual-color live- cell imaging experiments, we show that unencapsulated pneumococci adhere significantly better to human lung epithelial cells than encapsulated strains, in

Consistent with the epithelial cell analysis, we compared pneumococ- cal gene expression between unencapsulated and encapsulated librar- ies, resulting in 295 differentially

dari competence regulon pada studi berbasis array sebelumnya, sehingga penelitian yang telah kami lakukan memperlihatkan keunggulan teknologi sekuensing dalam riset