Reviews:

Microbial Cell, Vol. 5, No. 6, pp. 262 - 268; doi: 10.15698/mic2018.06.634

The CRISPR conundrum: evolve and maybe die, or survive and risk stagnation

Jesús García-Martínez1, Rafael D. Maldonado1, Noemí M. Guzmán1 and Francisco J. M. Mojica1,2

Download PDF download pdf
Show/hide additional information

    1 Departamento de Fisiología, Genética y Microbiología. Universidad de Alicante, Campus de San Vicente, 03690 San Vicente del Raspeig (Alicante), Spain.

    2 I.M.E.M. Ramón Margalef. Universidad de Alicante, Campus de San Vicente, 03690 San Vicente del Raspeig (Alicante), Spain.

Keywords: CRISPR, Cas, prokaryotic adaptive immunity, horizontal gene transfer, prokaryotic evolution.

Abbreviations:

crRNA – CRISPR RNA,

crRNP – CRISPR ribonucleoprotein,

PAM – proto-spacer adjacent motif,

tracrRNA – trans-activating crRNA.
Received originally: 26/03/2018 Received in revised form: 23/04/2018
Accepted: 25/04/2018 Published: 16/05/2018

Correspondence:
Francisco J. M. Mojica, Tel: +34 965909761. Fax +34 965909569; fmojica@ua.es

Conflict of interest statement: The authors have no conflict of interests to declare.
Please cite this article as: Jesús García-Martínez, Rafael D. Maldonado, Noemí M. Guzmán and Francisco J. M. Mojica (2018). The CRISPR conundrum: evolve and maybe die, or survive and risk stagnation. Microbial Cell 5(6): 262-268. doi: 10.15698/mic2018.06.634

Abstract

CRISPR-Cas represents a prokaryotic defense mechanism against invading genetic elements. Although there is a diversity of CRISPR-Cas systems, they all share similar, essential traits. In general, a CRISPR-Cas system consists of one or more groups of DNA repeats named CRISPR (Clustered Regularly Interspaced Short Palindromic Repeats), regularly separated by unique sequences referred to as spacers, and a set of functionally associated cas (CRISPR associated) genes typically located next to one of the repeat arrays. The origin of spacers is in many cases unknown but, when ascertained, they usually match foreign genetic molecules. The proteins encoded by some of the cas genes are in charge of the incorporation of new spacers upon entry of a genetic element. Other Cas proteins participate in generating CRISPR-spacer RNAs and perform the task of destroying nucleic acid molecules carrying sequences similar to the spacer. In this way, CRISPR-Cas provides protection against genetic intruders that could substantially affect the cell viability, thus acting as an adaptive immune system. However, this defensive action also hampers the acquisition of potentially beneficial, horizontally transferred genes, undermining evolution. Here we cover how the model bacterium Escherichia coli deals with CRISPR-Cas to tackle this major dilemma, evolution versus survival.

INTRODUCTION

The prokaryotic world has been historically the main source of tools for genetic engineering and molecular biology in general. CRISPR-Cas is a recent example of how the study of prokaryotes has revolutionized life sciences. Besides becoming the most important tool for genomic editing to date [1], the discovery of this immune system has marked an important milestone in the history of Microbiology.

Cas proteins, CRISPR loci and CRISPR RNAs are the core functional parts of an adaptive and heritable resistance system against foreign DNA. They enable the cell to keep memory of infections by exogenous elements and fight against the invader. There is a significant diversity of genes associated with CRISPR, presumably reflecting the selective pressure viruses exert on the evolution of the system. Classification of CRISPR-Cas systems has been proven a challenging task [2][3][4], and new variants are emerging as sequencing data increases and functional studies on these systems are performed. Distinct CRISPR-Cas systems can coexist in a genome [4][5][6][7][8]. Moreover, the number of CRISPR loci pertaining to the same type varies among organisms, and both the identity and number of spacers within each array greatly changes even among genomes of closely related strains [9].

In this paper, we present an overview of the CRISPR-Cas systems outlining their discovery, classification and functional role, and we discuss about the evolutionary importance of these systems in the model organism Escherichia coli. The chromosome of E. coli strains may harbor up to two CRISPR-Cas systems involving as much as two repeat arrays each [6]. Equivalent arrays show a considerable intraspecific polymorphism in terms of spacer number and sequence. Fundamental knowledge about the CRISPR-Cas mechanism has been generated from the analysis of these two systems in E. coli [10][11][12][13] and related Enterobacteriaceae [14].

DISCOVERY OF CRISPR LOCI AS DNA-MEMORY STORES

The serendipitous finding by Nakata and collaborators in 1987 [15] of five direct repeats next to the iap gene in E. coli was the first report of a CRISPR locus. Subsequently, in 1989 [16] the Nakata’s team documented another array of repeats at approximately 20 kb from the first one. Soon after, Hermans et al. [17] found direct repeats in the unrelated, Gram-positive Mycobacterium tuberculosis complex, launching the use of the repeat loci for strain typing based on their particular spacer content [18]. Archaea first CRISPRs were discovered in 1993 [19], and the earliest functional studies on these sequences were performed in 1995 [20]. By the end of the 1990’s, similar direct repeats were found in other prokaryotes and denominations given to these sequences started to multiply: DR, direct repeats [17]; TREPs, tandem repeats [20]; SRSR, short regularly spaced repeats [21]; DVR, direct variant repeats [22]; LCTR, large clusters of tandem repeats [23]; SPIDR, spacers interspersed direct repeats [24]. To avoid confusion, an agreement was made on naming the repeated sequences as CRISPR [25]. This acronym appeared published for the first time in 2002 [26]. By then, the biological relevance of these sequences was recognized, since they were distributed among many different, distantly related prokaryotes, representing a widespread family of repeats [21]. However, even though protein coding genes commonly associated to CRISPR arrays were discovered [26], unraveling their function was still pending. These Cas proteins, some of them related to helicases or nucleases, could play a role on DNA metabolism or expression[26].

Nevertheless, the definitive hint for the biological function of CRISPR-Cas came from the spacers rather than from the Cas or CRISPR units. In 2005, three independent studies found that some spacers matched sequences from transmissible genetic elements [27][28][29]. Notably, a comprehensive survey of the literature published on viruses and plasmids carrying spacer homologs, pointed to a relationship between immunity to these carriers and the presence of the cognate spacer in a potential host [27]. Therefore, it was suggested that the spacers represent a memory of past infections, and this information might be used to guide a defense mechanism. This fundamental breakthrough in the understanding of the CRISPR role in nature came hand in hand with the advent of increasing amounts of sequence data generated from viral, plasmid and complete genome sequences of prokaryotic strains which allowed researchers to cross-compare them. The existence of an adaptive, immunity-like system in Bacteria and Archaea was such an innovative idea that the three research groups undergone difficulties in publishing their results [30]. Historical perspectives of the initial moments of this discovery have been published elsewhere [25][30][31][32][33][34] showing interesting insights into the way modern science works and how scientific discoveries are made.

In 2007, the function of CRISPR-Cas as a specific immune system was experimentally proven in Streptococcus thermophilus [35]: phage resistance was endowed after the incorporation of small fragments of the foreign genetic material as spacers into the CRISPR loci of the bacterium. Moreover, Cas proteins were shown to be involved in this immunity. One year later it was demonstrated that transcripts derived from CRISPR arrays in E. coli were processed by Cas proteins and that the resulting small RNAs (crRNAs) are necessary to achieve immunity [12].

CRISPR-CAS MECHANISM

Despite the diversification of CRISPR-Cas systems and their wide distribution in distantly related bacteria and archaea [4], the fundamental mechanism of this immune system is quite conserved, following three basic steps: adaptation, expression and interference.

Adaptation, or spacer acquisition, requires the integration of fragments of nucleic acids from invader molecules [36][37][38]. In addition to Cas, non-Cas proteins are involved in this stage [39]. Fragments of foreign nucleic acids selected for integration, named proto-spacers [40], are usually flanked by short conserved sequences, the proto-spacer adjacent motif (PAM) [41]. New spacers are preferentially integrated in a polarized manner [29], next to the terminal CRISPR unit downstream to an AT-rich region called leader [12][26][42]. The PAM sequence is needed for most, but not all systems to recognize foreign targets, and its absence in the own CRISPR array avoids self-targeting [43]. Most CRISPR-Cas systems acquire spacers directly from DNA donors but a few systems are able to gain new spacers derived from RNA precursors after retrotranscription [44].

The transcription of a CRISPR array from the leader generates a multi-spacer RNA (pre-crRNA) which is processed to single-spacer crRNAs with the participation of Cas proteins [12] and, in some systems, of non-Cas ribonucleases as well as a trans-activating crRNA (tracrRNA) that partially hybridizes with the pre-crRNA [45]. After processing, each mature crRNA (or crRNA/tracrRNA duplex) remains assembled with Cas proteins in a CRISPR ribonucleoprotein (crRNP) complex [46][47]. This completes the second step of the CRISPR mechanism.

During the interference stage, the crRNP complex recognizes and directs cleavage of spacer-complementary sequences resulting in the elimination of molecules that carry potential targets [48]. Specific PAMs are crucial for efficient interference by many CRISPR-Cas systems [48][49][50]. In this case, upon PAM recognition by a protein of the crRNP complex, double-strand pairing is disrupted at the target DNA, leading to a R-loop conformation through progressive hybridization (starting from the PAM) with the spacer sequence in the crRNA [46]. The R-loop is the substrate for cleavage by Cas endonucleases [51]. Some CRISPR-Cas systems target RNA instead of, or in addition to, DNA [52][53][54].

CRISPR-Cas SYSTEMS CLASSIFICATION

Cas proteins are categorized in three functional modules [55]. The suite of proteins for the acquisition module is quite uniform. Regular members are Cas1 and Cas2 [36][56], which have nuclease activities and form a multi-protein complex [57]. The Cas1-Cas2 adaptation complex appears to be assisted by Cas4 when present, and might be included in this module [58][59]. In contrast to the acquisition proteins, the effector module (that is, proteins involved in pre-crRNA processing, target recognition and cleavage) is highly variable [3][4][60]. There is a third module of ancillary Cas proteins, involved in regulatory and other unknown roles [3][60].

Due to the fast evolution and wide diversification of the CRISPR-Cas systems, a multiple criteria approach has been used for classification: signature cas genes specific for some types, sequence similarity between common Cas proteins, the phylogeny of Cas1 (the most conserved Cas protein) and gene configuration in the loci [3][4]. The application of these criteria resulted in the current classification principle of two classes (1 and 2) and six types (from I to VI) [3]. Several subtypes (designated by letters, from ‘A’ forward) have been proposed based on signature genes and characteristic genomic arrangements [3][4]. Moreover, at least in the case of E. coli, subtype variants showing substantial differences in cas sequence and PAM preference have been recognized within the species [61][62]. This classification system also involves a systematic naming for Cas proteins that, in some cases, has changed over time to adapt to new discoveries [2][3][4].

Class 1 systems rely on multi-protein effector complexes [3]. They include Type I and Type III systems (distinguishable by the presence of Cas3 and Cas10, respectively) as well as the uncommon Type IV, devoid of an adaptation module. Class 2 is defined by the presence of a single-protein effector, namely Cas9, Cas12 or Cas13, depending on the particular type of system (Type II, Type V and Type VI, respectively) [3][63]. In spite of the need for tracrRNAs by Type II systems [45], not being involved in Class 1 systems [3], most applications of CRISPR technology in heterologous hosts are based on Type II components. This is mainly because, in contrast to Class 1, a single protein is required for interference and the target is cleaved just once at precise sites [31].

HOW PROKARYOTES BYPASS THE GENETIC BARRIER DICTATED BY CRISPR: THE CASE OF Escherichia coli

Once the biological function of CRISPR-Cas was revealed, the potential drawbacks that fully efficient CRISPR-mediated interference could pose to prokaryotic evolution became evident [64]. Horizontal Gene Transfer (HGT) is one of the main forces driving genetic change in Bacteria and Archaea [65][66]. However, the uptake of foreign nucleic acids might be constrained by functional CRISPR-Cas systems. To cope with such situation, prokaryotes either lack these systems or place them under stringent control [67][68]. This is exemplified by the case of E. coli, a paradigm of genome plasticity [69][70] in spite of being in possession of CRISPR-Cas systems [6]: a subtype I-E system is present in the majority of strains and a complete I-F system exists in a reduced number of isolates. Still, cells harboring cas genes of the two subtypes are extremely rare [6][61]. Unexpectedly, the E. coli I-F system is constitutively expressed under normal laboratory growth conditions [71][72]. Therefore, in principle, it is permanently acting against gene transfer. However, the PAMs of the spacers present in the I-F arrays of E. coli differ from the proto-spacer adjacent motifs that elicit the most efficient interference [71]. Such a relaxed interference could provide the opportunity for beneficial foreign DNA to be acquired, while at the same time still limiting exchange of unwanted genetic material. Remarkably, when I-F cas are absent in the E. coli genome, an array with a limited number of I-F repeats is invariably present, allegedly as a remnant of an ancient complete I-F system [6][61][73]. Most strikingly, the vast majority of spacers in these orphan arrays match cas I-F genes [6][73], playing a crucial role in preventing the barrier effect of their cognate genes [73]. Strains harboring these arrays use them as a constitutively expressed anti-cas mechanism that avoids the establishment of a fully equipped, immunity-prone CRISPR-Cas I-F system: intrusive DNA containing cas I-F genes is degraded through the action of the encoded Cas proteins guided by the resident crRNAs [73]. This anti-cas mechanism strongly supports the hypothesis that CRISPR-Cas immunity could be annoying for the carrier cell.

Opposite to I-F, expression of the E. coli I-E system is precisely regulated. H-NS protein is the main repressor of the system and its silencing effect can be lessened by the transcription factor LeuO [74][75][76]. The cAMP receptor protein (CRP) also contributes to CRISPR inhibition, acting as a competitor of LeuO for binding to the regulatory regions in the CRISPR-cas locus [77]. However, activity of the I-E system of E. coli has not been detected under the multiple laboratory growth conditions so far tested (our unpublished results), and the natural circumstances upon which such silencing is relieved remain to be clearly elucidated. In this regard, quorum sensing autoinducers of the N-Acyl-homoserine-lactone (AHL) class appear to activate CRISPR-Cas systems in Gram-negative bacteria such as Pseudomonas aeruginosa [78] and Serratia sp. [79] at elevated cell densities, when the risk of infection by bacteriophages is the highest [80]. Although this sort of induction has not been detected in pure cultures of E. coli, the presence in this species of AHL receptors [81][82] raises the possibility that their CRISPR-Cas systems might be regulated through an interspecific crosstalk, by signals secreted by other members of the microbial community. Overall, these findings illustrate the complexity of I-E CRISPR-Cas regulation in E. coli. Moreover, its diverging spacer count and identity within the species is an indication that CRISPR activity, at least at the adaptation stage, is turned on at a different pace depending on the particular group of strains.

Related to this, a notable case is that of pathogenic strains. When compared to non-pathogens (i.e., commensals), they gain a selective advantage via the acquisition through HGT of virulence factors, allowing them the ability to colonize more varied ecological niches within their hosts [70][83][84]. Inquiringly, a recent work from our group [85] established a negative correlation between pathogenicity and I-E CRISPR repeat count in E. coli: commensal strains tend to have more repeats than pathogenic isolates. This observation is compatible with the hypothesis that the activity of CRISPR-Cas I-E is kept limited when environmental adaptation needs to take precedence over protection, to minimize the negative effects of an evolutionary constraint. Another related question is why E. coli strains have lost either the I-E or the I-F cas genes, depending on their particular environment [6][85]. Indeed, most extra-intestinal pathogens pertaining to diverse phylogroups retain a I-F CRISPR-Cas system while the majority of commensals and enteric pathogens harbor a I-E system [85]. The preference for one or the other CRISPR-Cas subtype is suggestive of functional differences between the two systems. In this sense, previous works have reported that whereas spacers within I-E arrays of E. coli target viruses and plasmids alike, most I-F spacers matching known sequences have a plasmid origin [6][72][85]. Being plasmids the primary vectors for antibiotic resistance genes [86], this bias of I-F toward targeting plasmids is in agreement with the observation that the carrier strains are particularly susceptible to antibiotics [72]. Even though the reason for this apparent specialization is unknown, it highlights the inconvenience of an indiscriminate interference and the burden of carrying multiple CRISPR-Cas systems.

In summary, the analysis of the different CRISPR-Cas settings found in E. coli strengthen the idea that these systems, despite conferring protection, could severely hamper prokaryote evolution, hinting at how detrimental they could become if left unrestricted. Therefore, avoiding cas genes and limiting CRISPR-Cas activity when present appears to be a necessary evil for a prokaryote, where a delicate balance should be reached between the two extremes, those of promiscuity or chastity in terms of genetic exchange.

References

  1. M. Jinek, K. Chylinski, I. Fonfara, M. Hauer, J.A. Doudna, and E. Charpentier, "A Programmable Dual-RNA–Guided DNA Endonuclease in Adaptive Bacterial Immunity", Science, vol. 337, pp. 816-821, 2012. http://dx.doi.org/10.1126/science.1225829
  2. K.S. Makarova, D.H. Haft, R. Barrangou, S.J.J. Brouns, E. Charpentier, P. Horvath, S. Moineau, F.J.M. Mojica, Y.I. Wolf, A.F. Yakunin, J. van der Oost, and E.V. Koonin, "Evolution and classification of the CRISPR–Cas systems", Nature Reviews Microbiology, vol. 9, pp. 467-477, 2011. http://dx.doi.org/10.1038/nrmicro2577
  3. K.S. Makarova, Y.I. Wolf, O.S. Alkhnbashi, F. Costa, S.A. Shah, S.J. Saunders, R. Barrangou, S.J.J. Brouns, E. Charpentier, D.H. Haft, P. Horvath, S. Moineau, F.J.M. Mojica, R.M. Terns, M.P. Terns, M.F. White, A.F. Yakunin, R.A. Garrett, J. van der Oost, R. Backofen, and E.V. Koonin, "An updated evolutionary classification of CRISPR–Cas systems", Nature Reviews Microbiology, vol. 13, pp. 722-736, 2015. http://dx.doi.org/10.1038/nrmicro3569
  4. E.V. Koonin, K.S. Makarova, and F. Zhang, "Diversity, classification and evolution of CRISPR-Cas systems", Current Opinion in Microbiology, vol. 37, pp. 67-78, 2017. http://dx.doi.org/10.1016/j.mib.2017.05.008
  5. D.H. Haft, J. Selengut, E.F. Mongodin, and K.E. Nelson, "A Guild of 45 CRISPR-Associated (Cas) Protein Families and Multiple CRISPR/Cas Subtypes Exist in Prokaryotic Genomes", PLoS Computational Biology, vol. 1, pp. e60, 2005. http://dx.doi.org/10.1371/journal.pcbi.0010060
  6. C. Díez-Villaseñor, C. Almendros, J. García-Martínez, and F.J.M. Mojica, "Diversity of CRISPR loci in Escherichia coli", Microbiology, vol. 156, pp. 1351-1361, 2010. http://dx.doi.org/10.1099/mic.0.036046-0
  7. C. Rousseau, M. Gonnet, M. Le Romancer, and J. Nicolas, "CRISPI: a CRISPR interactive database", Bioinformatics, vol. 25, pp. 3317-3318, 2009. http://dx.doi.org/10.1093/bioinformatics/btp586
  8. Q. Zhang, and Y. Ye, "Not all predicted CRISPR–Cas systems are equal: isolated cas genes and classes of CRISPR like elements", BMC Bioinformatics, vol. 18, 2017. http://dx.doi.org/10.1186/s12859-017-1512-4
  9. I. Grissa, G. Vergnaud, and C. Pourcel, "The CRISPRdb database and tools to display CRISPRs and to generate dictionaries of spacers and repeats", BMC Bioinformatics, vol. 8, 2007. http://dx.doi.org/10.1186/1471-2105-8-172
  10. S. Redding, S. Sternberg, M. Marshall, B. Gibb, P. Bhat, C. Guegler, B. Wiedenheft, J. Doudna, and E. Greene, "Surveillance and Processing of Foreign DNA by the Escherichia coli CRISPR-Cas System", Cell, vol. 163, pp. 854-865, 2015. http://dx.doi.org/10.1016/j.cell.2015.10.003
  11. P. van Erp, R.N. Jackson, J. Carter, S.M. Golden, S. Bailey, and B. Wiedenheft, "Mechanism of CRISPR-RNA guided recognition of DNA targets inEscherichia coli", Nucleic Acids Research, vol. 43, pp. 8381-8391, 2015. http://dx.doi.org/10.1093/nar/gkv793
  12. S.J.J. Brouns, M.M. Jore, M. Lundgren, E.R. Westra, R.J.H. Slijkhuis, A.P.L. Snijders, M.J. Dickman, K.S. Makarova, E.V. Koonin, and J. van der Oost, "Small CRISPR RNAs Guide Antiviral Defense in Prokaryotes", Science, vol. 321, pp. 960-964, 2008. http://dx.doi.org/10.1126/science.1159689
  13. A. Levy, M.G. Goren, I. Yosef, O. Auster, M. Manor, G. Amitai, R. Edgar, U. Qimron, and R. Sorek, "CRISPR adaptation biases explain preference for acquisition of foreign DNA", Nature, vol. 520, pp. 505-510, 2015. http://dx.doi.org/10.1038/nature14302
  14. L. Medina-Aparicio, S. Dávila, J.E. Rebollar-Flores, E. Calva, and I. Hernández-Lucas, "The CRISPR-Cas system in Enterobacteriaceae", Pathogens and Disease, vol. 76, 2018. http://dx.doi.org/10.1093/femspd/fty002
  15. Y. Ishino, H. Shinagawa, K. Makino, M. Amemura, and A. Nakata, "Nucleotide sequence of the iap gene, responsible for alkaline phosphatase isozyme conversion in Escherichia coli, and identification of the gene product", Journal of Bacteriology, vol. 169, pp. 5429-5433, 1987. http://dx.doi.org/10.1128/jb.169.12.5429-5433.1987
  16. A. Nakata, M. Amemura, and K. Makino, "Unusual nucleotide arrangement with repeated sequences in the Escherichia coli K-12 chromosome", Journal of Bacteriology, vol. 171, pp. 3553-3556, 1989. http://dx.doi.org/10.1128/jb.171.6.3553-3556.1989
  17. P.W. Hermans, D. van Soolingen, E.M. Bik, P.E. de Haas, J.W. Dale, and J.D. van Embden, "Insertion element IS987 from Mycobacterium bovis BCG is located in a hot-spot integration region for insertion elements in Mycobacterium tuberculosis complex strains.", Infection and immunity, 1991. http://www.ncbi.nlm.nih.gov/pubmed/1649798
  18. P.M.A. Groenen, A.E. Bunschoten, D.V. Soolingen, and J.D.A.V. Errtbden, "Nature of DNA polymorphism in the direct repeat cluster of Mycobacterium tuberculosis; application for strain differentiation by a novel typing method", Molecular Microbiology, vol. 10, pp. 1057-1065, 1993. http://dx.doi.org/10.1111/j.1365-2958.1993.tb00976.x
  19. F.J.M. Mojica, G. Juez, and F. Rodriguez‐Valera, "Transcription at different salinities of Haloferax mediterranei sequences adjacent to partially modified PstI sites", Molecular Microbiology, vol. 9, pp. 613-621, 1993. http://dx.doi.org/10.1111/j.1365-2958.1993.tb01721.x
  20. F. Mojica, C. Ferrer, G. Juez, and F. Rodríguez‐Valera, "Long stretches of short tandem repeats are present in the largest replicons of the Archaea Haloferax mediterranei and Haloferax volcanii and could be involved in replicon partitioning", Molecular Microbiology, vol. 17, pp. 85-93, 1995. http://dx.doi.org/10.1111/j.1365-2958.1995.mmi_17010085.x
  21. F.J.M. Mojica, C. Díez‐Villaseñor, E. Soria, and G. Juez, "Biological significance of a family of regularly spaced repeats in the genomes of Archaea, Bacteria and mitochondria", Molecular Microbiology, vol. 36, pp. 244-246, 2000. http://dx.doi.org/10.1046/j.1365-2958.2000.01838.x
  22. J.D.A. van Embden, T. van Gorkom, K. Kremer, R. Jansen, B.A.M. van der Zeijst, and L.M. Schouls, "Genetic Variation and Evolutionary Origin of the Direct Repeat Locus of Mycobacterium tuberculosis Complex Bacteria", Journal of Bacteriology, vol. 182, pp. 2393-2401, 2000. http://dx.doi.org/10.1128/JB.182.9.2393-2401.2000
  23. Q. She, R.K. Singh, F. Confalonieri, Y. Zivanovic, G. Allard, M.J. Awayez, C.C. Chan-Weiher, I.G. Clausen, B.A. Curtis, A. De Moors, G. Erauso, C. Fletcher, P.M.K. Gordon, I. Heikamp-de Jong, A.C. Jeffries, C.J. Kozera, N. Medina, X. Peng, H.P. Thi-Ngoc, P. Redder, M.E. Schenk, C. Theriault, N. Tolstrup, R.L. Charlebois, W.F. Doolittle, M. Duguet, T. Gaasterland, R.A. Garrett, M.A. Ragan, C.W. Sensen, and J. Van der Oost, "The complete genome of the crenarchaeon Sulfolobus solfataricus P2", Proceedings of the National Academy of Sciences, vol. 98, pp. 7835-7840, 2001. http://dx.doi.org/10.1073/pnas.141222098
  24. R. Jansen, J.D. van Embden, W. Gaastra, and L.M. Schouls, "Identification of a Novel Family of Sequence Repeats among Prokaryotes", OMICS: A Journal of Integrative Biology, vol. 6, pp. 23-33, 2002. http://dx.doi.org/10.1089/15362310252780816
  25. F. Mojica, and R. Garrett, " Discovery and seminal developments in the CRISPR field. In: Barrangou R, van der Oost J, editors. CRISPR-Cas systems: RNA-mediated adaptive immunity in Bacteria and Archaea.", Springer, Berlin, Heidelberg; pp 1–31., 2013.
  26. R. Jansen, J.D.A.V. Embden, W. Gaastra, and L.M. Schouls, "Identification of genes that are associated with DNA repeats in prokaryotes", Molecular Microbiology, vol. 43, pp. 1565-1575, 2002. http://dx.doi.org/10.1046/j.1365-2958.2002.02839.x
  27. F.J. Mojica, C. D�ez-Villase�or, J. Garc�a-Mart�nez, and E. Soria, "Intervening Sequences of Regularly Spaced Prokaryotic Repeats Derive from Foreign Genetic Elements", Journal of Molecular Evolution, vol. 60, pp. 174-182, 2005. http://dx.doi.org/10.1007/s00239-004-0046-3
  28. A. Bolotin, B. Quinquis, A. Sorokin, and S.D. Ehrlich, "Clustered regularly interspaced short palindrome repeats (CRISPRs) have spacers of extrachromosomal origin", Microbiology, vol. 151, pp. 2551-2561, 2005. http://dx.doi.org/10.1099/mic.0.28048-0
  29. C. Pourcel, G. Salvignol, and G. Vergnaud, "CRISPR elements in Yersinia pestis acquire new repeats by preferential uptake of bacteriophage DNA, and provide additional tools for evolutionary studies", Microbiology, vol. 151, pp. 653-663, 2005. http://dx.doi.org/10.1099/mic.0.27437-0
  30. E. Lander, "The Heroes of CRISPR", Cell, vol. 164, pp. 18-28, 2016. http://dx.doi.org/10.1016/j.cell.2015.12.041
  31. F.J. Mojica, and L. Montoliu, "On the Origin of CRISPR-Cas Technology: From Prokaryotes to Mammals", Trends in Microbiology, vol. 24, pp. 811-820, 2016. http://dx.doi.org/10.1016/j.tim.2016.06.005
  32. F.J.M. Mojica, and F. Rodriguez‐Valera, "The discovery of CRISPR in archaea and bacteria", The FEBS Journal, vol. 283, pp. 3162-3169, 2016. http://dx.doi.org/10.1111/febs.13766
  33. R. Barrangou, and P. Horvath, "A decade of discovery: CRISPR functions and applications", Nature Microbiology, vol. 2, 2017. http://dx.doi.org/10.1038/nmicrobiol.2017.92
  34. M. Morange, "What history tells us XXXVII. CRISPR-Cas: The discovery of an immune system in prokaryotes", Journal of Biosciences, vol. 40, pp. 221-223, 2015. http://dx.doi.org/10.1007/s12038-015-9532-6
  35. R. Barrangou, C. Fremaux, H. Deveau, M. Richards, P. Boyaval, S. Moineau, D.A. Romero, and P. Horvath, "CRISPR Provides Acquired Resistance Against Viruses in Prokaryotes", Science, vol. 315, pp. 1709-1712, 2007. http://dx.doi.org/10.1126/science.1138140
  36. G. Amitai, and R. Sorek, "CRISPR–Cas adaptation: insights into the mechanism of action", Nature Reviews Microbiology, vol. 14, pp. 67-76, 2016. http://dx.doi.org/10.1038/nrmicro.2015.14
  37. S.A. Jackson, R.E. McKenzie, R.D. Fagerlund, S.N. Kieper, P.C. Fineran, and S.J.J. Brouns, "CRISPR-Cas: Adapting to change", Science, vol. 356, 2017. http://dx.doi.org/10.1126/science.aal5056
  38. S. Sternberg, H. Richter, E. Charpentier, and U. Qimron, "Adaptation in CRISPR-Cas Systems", Molecular Cell, vol. 61, pp. 797-808, 2016. http://dx.doi.org/10.1016/j.molcel.2016.01.030
  39. T. Killelea, and E. Bolt, "CRISPR-Cas adaptive immunity and the three Rs", Bioscience Reports, vol. 37, 2017. http://dx.doi.org/10.1042/BSR20160297
  40. H. Deveau, R. Barrangou, J.E. Garneau, J. Labonté, C. Fremaux, P. Boyaval, D.A. Romero, P. Horvath, and S. Moineau, "Phage Response to CRISPR-Encoded Resistance in Streptococcus thermophilus", Journal of Bacteriology, vol. 190, pp. 1390-1400, 2008. http://dx.doi.org/10.1128/JB.01412-07
  41. F.J.M. Mojica, C. Díez-Villaseñor, J. García-Martínez, and C. Almendros, "Short motif sequences determine the targets of the prokaryotic CRISPR defence system", Microbiology, vol. 155, pp. 733-740, 2009. http://dx.doi.org/10.1099/mic.0.023960-0
  42. R. Lillestøl, P. Redder, R.A. Garrett, and K. Brügger, "A putative viral defence mechanism in archaeal cells", Archaea, vol. 2, pp. 59-72, 2006. http://dx.doi.org/10.1155/2006/542818
  43. E.R. Westra, E. Semenova, K.A. Datsenko, R.N. Jackson, B. Wiedenheft, K. Severinov, and S.J.J. Brouns, "Type I-E CRISPR-Cas Systems Discriminate Target from Non-Target DNA through Base Pairing-Independent PAM Recognition", PLoS Genetics, vol. 9, pp. e1003742, 2013. http://dx.doi.org/10.1371/journal.pgen.1003742
  44. S. Silas, G. Mohr, D.J. Sidote, L.M. Markham, A. Sanchez-Amat, D. Bhaya, A.M. Lambowitz, and A.Z. Fire, "Direct CRISPR spacer acquisition from RNA by a natural reverse transcriptase–Cas1 fusion protein", Science, vol. 351, 2016. http://dx.doi.org/10.1126/science.aad4234
  45. E. Deltcheva, K. Chylinski, C.M. Sharma, K. Gonzales, Y. Chao, Z.A. Pirzada, M.R. Eckert, J. Vogel, and E. Charpentier, "CRISPR RNA maturation by trans-encoded small RNA and host factor RNase III", Nature, vol. 471, pp. 602-607, 2011. http://dx.doi.org/10.1038/nature09886
  46. M.M. Jore, M. Lundgren, E. van Duijn, J.B. Bultema, E.R. Westra, S.P. Waghmare, B. Wiedenheft, . Pul, R. Wurm, R. Wagner, M.R. Beijer, A. Barendregt, K. Zhou, A.P.L. Snijders, M.J. Dickman, J.A. Doudna, E.J. Boekema, A.J.R. Heck, J. van der Oost, and S.J.J. Brouns, "Structural basis for CRISPR RNA-guided DNA recognition by Cascade", Nature Structural & Molecular Biology, vol. 18, pp. 529-536, 2011. http://dx.doi.org/10.1038/nsmb.2019
  47. C.R. Hale, P. Zhao, S. Olson, M.O. Duff, B.R. Graveley, L. Wells, R.M. Terns, and M.P. Terns, "RNA-Guided RNA Cleavage by a CRISPR RNA-Cas Protein Complex", Cell, vol. 139, pp. 945-956, 2009. http://dx.doi.org/10.1016/j.cell.2009.07.040
  48. J.E. Garneau, M. Dupuis, M. Villion, D.A. Romero, R. Barrangou, P. Boyaval, C. Fremaux, P. Horvath, A.H. Magadán, and S. Moineau, "The CRISPR/Cas bacterial immune system cleaves bacteriophage and plasmid DNA", Nature, vol. 468, pp. 67-71, 2010. http://dx.doi.org/10.1038/nature09523
  49. B. Zetsche, J. Gootenberg, O. Abudayyeh, I. Slaymaker, K. Makarova, P. Essletzbichler, S. Volz, J. Joung, J. van der Oost, A. Regev, E. Koonin, and F. Zhang, "Cpf1 Is a Single RNA-Guided Endonuclease of a Class 2 CRISPR-Cas System", Cell, vol. 163, pp. 759-771, 2015. http://dx.doi.org/10.1016/j.cell.2015.09.038
  50. E. Semenova, M.M. Jore, K.A. Datsenko, A. Semenova, E.R. Westra, B. Wanner, J. van der Oost, S.J.J. Brouns, and K. Severinov, "Interference by clustered regularly interspaced short palindromic repeat (CRISPR) RNA is governed by a seed sequence", Proceedings of the National Academy of Sciences, vol. 108, pp. 10098-10103, 2011. http://dx.doi.org/10.1073/pnas.1104144108
  51. S.H. Sternberg, S. Redding, M. Jinek, E.C. Greene, and J.A. Doudna, "DNA interrogation by the CRISPR RNA-guided endonuclease Cas9", Nature, vol. 507, pp. 62-67, 2014. http://dx.doi.org/10.1038/nature13011
  52. A. Hatoum-Aslan, K. Palmer, M. Gilmore, and L. Marraffini, "Type III CRISPR-Cas systems and the roles of CRISPR-Cas in bacterial virulence. In: Barrangou R, van der Oost J, editors. CRISPR-Cas systems: RNA-mediated adaptive immunity in Bacteria and Archaea.", Springer, Berlin, Heidelberg; pp 201–219., 2013.
  53. O.O. Abudayyeh, J.S. Gootenberg, P. Essletzbichler, S. Han, J. Joung, J.J. Belanto, V. Verdine, D.B.T. Cox, M.J. Kellner, A. Regev, E.S. Lander, D.F. Voytas, A.Y. Ting, and F. Zhang, "RNA targeting with CRISPR–Cas13", Nature, vol. 550, pp. 280-284, 2017. http://dx.doi.org/10.1038/nature24049
  54. B.A. Rousseau, Z. Hou, M.J. Gramelspacher, and Y. Zhang, "Programmable RNA Cleavage and Recognition by a Natural CRISPR-Cas9 System from Neisseria meningitidis", Molecular Cell, vol. 69, pp. 906-914.e4, 2018. http://dx.doi.org/10.1016/j.molcel.2018.01.025
  55. K. Makarova, Y. Wolf, and E. Koonin, "The basic building blocks and evolution of CRISPR–Cas systems", Biochemical Society Transactions, vol. 41, pp. 1392-1400, 2013. http://dx.doi.org/10.1042/BST20130038
  56. I. Yosef, M.G. Goren, and U. Qimron, "Proteins and DNA elements essential for the CRISPR adaptation process in Escherichia coli", Nucleic Acids Research, vol. 40, pp. 5569-5576, 2012. http://dx.doi.org/10.1093/nar/gks216
  57. J.K. Nuñez, P.J. Kranzusch, J. Noeske, A.V. Wright, C.W. Davies, and J.A. Doudna, "Cas1–Cas2 complex formation mediates spacer acquisition during CRISPR–Cas adaptive immunity", Nature Structural & Molecular Biology, vol. 21, pp. 528-534, 2014. http://dx.doi.org/10.1038/nsmb.2820
  58. S.N. Kieper, C. Almendros, J. Behler, R.E. McKenzie, F.L. Nobrega, A.C. Haagsma, J.N. Vink, W.R. Hess, and S.J. Brouns, "Cas4 Facilitates PAM-Compatible Spacer Selection during CRISPR Adaptation", Cell Reports, vol. 22, pp. 3377-3384, 2018. http://dx.doi.org/10.1016/j.celrep.2018.02.103
  59. H. Lee, Y. Zhou, D.W. Taylor, and D.G. Sashital, "Cas4-Dependent Prespacer Processing Ensures High-Fidelity Programming of CRISPR Arrays", Molecular Cell, vol. 70, pp. 48-59.e5, 2018. http://dx.doi.org/10.1016/j.molcel.2018.03.003
  60. P. Mohanraju, K.S. Makarova, B. Zetsche, F. Zhang, E.V. Koonin, and J. van der Oost, "Diverse evolutionary roots and mechanistic variations of the CRISPR-Cas systems", Science, vol. 353, 2016. http://dx.doi.org/10.1126/science.aad5147
  61. C. Almendros, F.J.M. Mojica, C. Díez-Villaseñor, N.M. Guzmán, and J. García-Martínez, "CRISPR-Cas Functional Module Exchange in Escherichia coli", mBio, vol. 5, 2014. http://dx.doi.org/10.1128/mBio.00767-13
  62. C. Díez-Villaseñor, N.M. Guzmán, C. Almendros, J. García-Martínez, and F.J. Mojica, "CRISPR-spacer integration reporter plasmids reveal distinct genuine acquisition specificities among CRISPR-Cas I-E variants ofEscherichia coli", RNA Biology, vol. 10, pp. 792-802, 2013. http://dx.doi.org/10.4161/rna.24023
  63. S. Shmakov, A. Smargon, D. Scott, D. Cox, N. Pyzocha, W. Yan, O.O. Abudayyeh, J.S. Gootenberg, K.S. Makarova, Y.I. Wolf, K. Severinov, F. Zhang, and E.V. Koonin, "Diversity and evolution of class 2 CRISPR–Cas systems", Nature Reviews Microbiology, vol. 15, pp. 169-182, 2017. http://dx.doi.org/10.1038/nrmicro.2016.184
  64. L.A. Marraffini, and E.J. Sontheimer, "CRISPR Interference Limits Horizontal Gene Transfer in Staphylococci by Targeting DNA", Science, vol. 322, pp. 1843-1845, 2008. http://dx.doi.org/10.1126/science.1165771
  65. L.D. McDaniel, E. Young, J. Delaney, F. Ruhnau, K.B. Ritchie, and J.H. Paul, "High Frequency of Horizontal Gene Transfer in the Oceans", Science, vol. 330, pp. 50-50, 2010. http://dx.doi.org/10.1126/science.1192243
  66. C.M. Thomas, and K.M. Nielsen, "Mechanisms of, and Barriers to, Horizontal Gene Transfer between Bacteria", Nature Reviews Microbiology, vol. 3, pp. 711-721, 2005. http://dx.doi.org/10.1038/nrmicro1234
  67. A.G. Patterson, M.S. Yevstigneyeva, and P.C. Fineran, "Regulation of CRISPR–Cas adaptive immune systems", Current Opinion in Microbiology, vol. 37, pp. 1-7, 2017. http://dx.doi.org/10.1016/j.mib.2017.02.004
  68. L.M. Leon, S.D. Mendoza, and J. Bondy-Denomy, "How bacteria control the CRISPR-Cas arsenal", Current Opinion in Microbiology, vol. 42, pp. 87-95, 2018. http://dx.doi.org/10.1016/j.mib.2017.11.005
  69. A. Porse, H. Gumpert, J.Z. Kubicek-Sutherland, N. Karami, I. Adlerberth, A.E. Wold, D.I. Andersson, and M.O.A. Sommer, "Genome Dynamics of Escherichia coli during Antibiotic Treatment: Transfer, Loss, and Persistence of Genetic Elements In situ of the Infant Gut", Frontiers in Cellular and Infection Microbiology, vol. 7, 2017. http://dx.doi.org/10.3389/fcimb.2017.00126
  70. U. Dobrindt, "(Patho-)Genomics of Escherichia coli", International Journal of Medical Microbiology, vol. 295, pp. 357-371, 2005. http://dx.doi.org/10.1016/j.ijmm.2005.07.009
  71. C. Almendros, N.M. Guzmán, C. Díez-Villaseñor, J. García-Martínez, and F.J.M. Mojica, "Target Motifs Affecting Natural Immunity by a Constitutive CRISPR-Cas System in Escherichia coli", PLoS ONE, vol. 7, pp. e50797, 2012. http://dx.doi.org/10.1371/journal.pone.0050797
  72. S. Aydin, Y. Personne, E. Newire, R. Laverick, O. Russell, A.P. Roberts, and V.I. Enne, "Presence of Type I-F CRISPR/Cas systems is associated with antimicrobial susceptibility in Escherichia coli", Journal of Antimicrobial Chemotherapy, vol. 72, pp. 2213-2218, 2017. http://dx.doi.org/10.1093/jac/dkx137
  73. C. Almendros, N.M. Guzmán, J. García-Martínez, and F.J.M. Mojica, "Anti-cas spacers in orphan CRISPR4 arrays prevent uptake of active CRISPR–Cas I-F systems", Nature Microbiology, vol. 1, 2016. http://dx.doi.org/10.1038/nmicrobiol.2016.81
  74. . Pul, R. Wurm, Z. Arslan, R. Geißen, N. Hofmann, and R. Wagner, "Identification and characterization ofE. coliCRISPR-caspromoters and their silencing by H-NS", Molecular Microbiology, vol. 75, pp. 1495-1512, 2010. http://dx.doi.org/10.1111/j.1365-2958.2010.07073.x
  75. E.R. Westra, . Pul, N. Heidrich, M.M. Jore, M. Lundgren, T. Stratmann, R. Wurm, A. Raine, M. Mescher, L. Van Heereveld, M. Mastop, E.G.H. Wagner, K. Schnetz, J. Van Der Oost, R. Wagner, and S.J.J. Brouns, "H‐NS‐mediated repression of CRISPR‐based immunity in Escherichia coli K12 can be relieved by the transcription activator LeuO", Molecular Microbiology, vol. 77, pp. 1380-1393, 2010. http://dx.doi.org/10.1111/j.1365-2958.2010.07315.x
  76. K. Pougach, E. Semenova, E. Bogdanova, K.A. Datsenko, M. Djordjevic, B.L. Wanner, and K. Severinov, "Transcription, processing and function of CRISPR cassettes in Escherichia coli", Molecular Microbiology, vol. 77, pp. 1367-1379, 2010. http://dx.doi.org/10.1111/j.1365-2958.2010.07265.x
  77. C. Yang, Y. Chen, H. Huang, H. Huang, and C. Tseng, "CRP represses the CRISPR/Cas system in Escherichia coli: evidence that endogenous CRISPR spacers impede phage P1 replication", Molecular Microbiology, vol. 92, pp. 1072-1091, 2014. http://dx.doi.org/10.1111/mmi.12614
  78. N.M. Høyland-Kroghsbo, J. Paczkowski, S. Mukherjee, J. Broniewski, E. Westra, J. Bondy-Denomy, and B.L. Bassler, "Quorum sensing controls the Pseudomonas aeruginosa CRISPR-Cas adaptive immune system", Proceedings of the National Academy of Sciences, vol. 114, pp. 131-135, 2016. http://dx.doi.org/10.1073/pnas.1617415113
  79. A.G. Patterson, S.A. Jackson, C. Taylor, G.B. Evans, G.P. Salmond, R. Przybilski, R.H. Staals, and P.C. Fineran, "Quorum Sensing Controls Adaptive Immunity through the Regulation of Multiple CRISPR-Cas Systems", Molecular Cell, vol. 64, pp. 1102-1108, 2016. http://dx.doi.org/10.1016/j.molcel.2016.11.012
  80. B. Knowles, C.B. Silveira, B.A. Bailey, K. Barott, V.A. Cantu, A.G. Cobián-Güemes, F.H. Coutinho, E.A. Dinsdale, B. Felts, K.A. Furby, E.E. George, K.T. Green, G.B. Gregoracci, A.F. Haas, J.M. Haggerty, E.R. Hester, N. Hisakawa, L.W. Kelly, Y.W. Lim, M. Little, A. Luque, T. McDole-Somera, K. McNair, L.S. de Oliveira, S.D. Quistad, N.L. Robinett, E. Sala, P. Salamon, S.E. Sanchez, S. Sandin, G.G.Z. Silva, J. Smith, C. Sullivan, C. Thompson, M.J.A. Vermeij, M. Youle, C. Young, B. Zgliczynski, R. Brainard, R.A. Edwards, J. Nulton, F. Thompson, and F. Rohwer, "Lytic to temperate switching of viral communities", Nature, vol. 531, pp. 466-470, 2016. http://dx.doi.org/10.1038/nature17193
  81. J. Lee, T. Maeda, S.H. Hong, and T.K. Wood, "Reconfiguring the Quorum-Sensing Regulator SdiA ofEscherichia coliTo Control Biofilm Formation via Indole andN-Acylhomoserine Lactones", Applied and Environmental Microbiology, vol. 75, pp. 1703-1716, 2009. http://dx.doi.org/10.1128/AEM.02081-08
  82. V. Sperandio, "SdiA sensing of acyl-homoserine lactones by enterohemorrhagicE. coli(EHEC) serotype O157", Gut Microbes, vol. 1, pp. 432-435, 2010. http://dx.doi.org/10.4161/gmic.1.6.14177
  83. J.B. Kaper, J.P. Nataro, and H.L.T. Mobley, "Pathogenic Escherichia coli", Nature Reviews Microbiology, vol. 2, pp. 123-140, 2004. http://dx.doi.org/10.1038/nrmicro818
  84. M. Diard, and W. Hardt, "Evolution of bacterial virulence", FEMS Microbiology Reviews, vol. 41, pp. 679-697, 2017. http://dx.doi.org/10.1093/femsre/fux023
  85. E. García-Gutiérrez, C. Almendros, F.J.M. Mojica, N.M. Guzmán, and J. García-Martínez, "CRISPR Content Correlates with the Pathogenic Potential of Escherichia coli", PLOS ONE, vol. 10, pp. e0131935, 2015. http://dx.doi.org/10.1371/journal.pone.0131935
  86. J.M.A. Blair, M.A. Webber, A.J. Baylay, D.O. Ogbolu, and L.J.V. Piddock, "Molecular mechanisms of antibiotic resistance", Nature Reviews Microbiology, vol. 13, pp. 42-51, 2014. http://dx.doi.org/10.1038/nrmicro3380

ACKNOWLEDGMENTS

The authors are supported by grants BIO2014-53029-P (Ministerio de Economía y Competitividad, Spain), 291815 Era-Net ANIHWA (7th Framework Programme, European Commission) and PROMETEO/2017/129 (Conselleria d'Educació, Investigació, Cultura i Esport, Generalitat Valenciana, Spain).

COPYRIGHT

© 2018

Creative Commons License
The CRISPR conundrum: evolve and maybe die, or survive and risk stagnation by García-Martínez et al. is licensed under a Creative Commons Attribution 4.0 International License.

By continuing to use the site, you agree to the use of cookies. more information

The cookie settings on this website are set to "allow cookies" to give you the best browsing experience possible. If you continue to use this website without changing your cookie settings or you click "Accept" below then you are consenting to this. Please refer to our "privacy statement" and our "terms of use" for further information.

Close