• No results found

Requirements for the targeting of foreignDNA by the Escherichia coli CRISPR/CassystemMirthe Hoekzema

N/A
N/A
Protected

Academic year: 2022

Share "Requirements for the targeting of foreignDNA by the Escherichia coli CRISPR/CassystemMirthe Hoekzema"

Copied!
17
0
0

Loading.... (view fulltext now)

Full text

(1)

Requirements for the targeting of foreign DNA by the Escherichia coli CRISPR/Cas system

Mirthe Hoekzema

Degree project in applied biotechnology, Master of Science (2 years), 2011 Examensarbete i tillämpad bioteknik 45 hp till masterexamen, 2011

Biology Education Centre and Department of Cell and Molecular Biology, Uppsala University

(2)
(3)

The CRISPR/Cas system is a recently discovered adaptive prokaryotic immune system against invaders such as phages and plasmids. The system consists of one or more Clustered Regularly Interspaced Short Palindromic Repeats (CRISPR) loci that are transcribed from a leader region. The primary transcript is processed by the CRISPR associated (Cas) proteins into small crRNAs that guide the defense apparatus and direct cleavage of the target nucleic acid. Specificity is achieved by base pairing of the extrachromosomally derived spacer sequences of the crRNAs and the complementary sequence in the target, called proto-spacer, flanked by a di-or trinucleotide motif named PAM (for proto-spacer adjacent motif). CRISPR/Cas systems are highly variable and widely spread throughout prokaryotes. Here we investigate the CRISPR/Cas system of the commonly used model organism Escherichia coli K12. Since no CRISPR activity has been

documented in the wild type E. coli K12, likely due to low levels of Cas proteins, we have modified the E. coli chromosome to overexpress these proteins by placing a promoter in front of the cas operon. By introducing a synthetic CRISPR- array the system was directed to target the coli phage lambda. Using these

modified strains we hope to further elucidate the CRISPR/Cas interference mechanism. So far we have found that an upstream PAM, or PAM-like motif, is needed for targeting of invading plasmids. Ongoing investigations focus on visualizing the site-specific cleavage of target DNA.

Introduction

Already in 1987 Ishino and co-workers noticed the peculiar direct repeat sequences interspaced by variable but regularly sized spacer regions downstream of the iap gene in E. coli (Ishino et al. 1987), and subsequent discoveries of similar repetitive motifs in other bacterial and archaeal species were made (Groenen et al. 1993, Mojica et al. 1995, She et al.

2001), but it took almost two decades before its biological function as an adaptive immune system became evident. The repetitive motifs were termed CRISPR for Clustered Regularly Interspaced Short Palindromic Repeats (Jansen et al. 2002). The breakthrough in understanding CRISPR function came in 2005 when three independent research groups reported that the spacer regions between the repeats often had an extrachromosomal origin, being homologous to sequences found in phage and plasmid DNA (Bolotin et al. 2005, Mojica et al. 2005, Pourcel et al. 2005). This implied a function as a specific immune system, which was firmly established by Barrangou and co-workers in 2007 when they provided in vivo evidence that the presence of a spacer matching a phage sequence, together with Cas proteins, provides resistance against the phage containing the particular sequence (Barrangou et al. 2007). Since then our understanding of the CRISPR/Cas system has expanded and a complex picture has emerged.

CRISPR arrays consist of 26 to 72 bp equally sized but variable sequences called spacers, that are flanked by 21 to 48 bp direct repeats. As many as 274 or as few as 1

spacer/repeat unit per CRISPR locus have been found, the current average being 66

(Deveau et al. 2010, Marraffini and Sontheimer 2010). Twelve categories of repeat

sequences can be distinguished based on sequence similarity and secondary structure

(Kunin et al. 2007). The repeats are identical within each locus but can be different

between closely related species or even between different loci within a given species. The

consensus is that the majority of spacers are sequences derived from phage genomes or

plasmids (Bolotin et al. 2005, Mojica et al. 2005, Pourcel et al. 2005) but occasionally

spacers targeting endogenous genes are found (Stern et al. 2010). The sequence in the

target DNA corresponding to the spacer is called proto-spacer, immediately next to the

proto-spacer is the proto-spacer adjacent motif or PAM (Deveau et al. 2008, Horvath et

(4)

al. 2008, Mojica et al. 2009). PAMs are CRISPR type specific and likely to be involved in spacer acquisition (Mojica et al. 2009), as well as CRISPR interference, as phages have been shown to be able to avoid targeting by the CRISPR/Cas system by mutating the PAM motif (Deveau et al. 2008). Upstream of the CRISPR array is the A/T rich leader sequence (Jansen et al. 2002). In E. coli it has been shown that this sequence contains the promoter driving transcription of the CRISPR locus (Pul et al. 2010) that seems to be subject to H-NS silencing (Pul et al. 2010, Westra et al. 2010). Cas (CRISPR associated) genes are found exclusively in genomes bearing CRISPR loci, often in their close proximity, and they typically have domains characteristic for helicases, nucleases and DNA binding proteins (Jansen et al. 2002, Haft et al. 2005). The CRISPR array, the leader sequences and the Cas proteins together comprise the CRISPR/Cas system.  

The mechanism behind CRISPR/Cas mediated immunity is not yet completely

understood. The existing working model (Figure 1) distinguishes three stages, adaptation, expression and interference (van der Oost et al. 2009). The adaptation stage is when immunization happens, a new genomic invader is encountered and by an unknown mechanism, a piece of the foreign DNA is integrated at the leader end of the CRISPR array (Barrangou et al. 2007, Deveau et al. 2008, Horvath et al. 2008, Tyson and Banfield 2008). The conserved Cas1 and Cas2 proteins are likely players in this process (Brouns et al. 2008). How the foreign DNA is recognized is not known. With the spacer a new repeat is added as well, most likely by duplication of the adjacent repeat (van der Oost et al. 2009). The second stage is expression of the CRISPR array from the leader, and processing of the primary transcript by endoribonucleases, generating small crRNAs. In E. coli the Cascade complex processes the pre-crRNA, and the crRNAs remain associated with the complex (Brouns et al. 2008). The crRNAs contains one complete spacer

sequence typically flanked by 8 nucleotides of the repeat at the 5’ end, while the length of the 3’ end varies between species (Al-attar et al. 2011, Makarova et al. 2011). The cr-

cas3 Pcas A B C D E 1 2

L

Interference Expression

Cas3

Cascade

Transcription pre-crRNA Processing

(Cascade)

crRNAs Cascade-crRNA

Targeting Phage dsDNA

AUAAACCGUGGGAUGCCUACCGCAAGCAGCUUGGCCUGAAGAGUUCCCC |||||||||||||||||||||||||||||||||

--AAGGGCATACACCCTACGGATGGCGTTCGTCGAACCGGACTTTCTGAAGAGAGGC-- --TTCCCGTATGTGGGATGCCTACCGCAAGCAGCTTGGCCTGAAAGACTTCTCTCCG--

PAM

L

Foreign DNA Integration

CRISPR locus

Adaptation

Figure 1; The CRISPR/Cas system of E. coli

(5)

RNAs guide an interference complex to the foreign nucleic acid, which subsequently eliminates the target by cleavage within the proto-spacer (Garneau et al. 2010) in what is called the interference stage. In E. coli, Cas3 is the most likely catalyst of the cleavage reaction, as it has been shown to be essential for interference but not for generation of the crRNAs (Brouns et al. 2008). DNA, not RNA, is the likely target for most

CRISPR/Cas systems (Makarova et al. 2011). A recent study in E. coli showed Cascade bound crRNA form Watson-Crick base pairs with the complementary strand in double stranded DNA forming a R-loop (Jore et al. 2011), strongly suggesting that the

CRISPR/cas system of E. coli operates by cleavage of foreign DNA.

The CRISPR/Cas systems are found in about half of the bacterial and nearly all archaeal chromosomes sequenced (Grissa et al. 2007), and recently a new classification system was proposed (Makarova et al. 2011) This system divides the CRISPR/Cas systems into three major types with a further division into subtypes based on the phylogeny of the common cas genes, CRISPR repeat sequence, and overall architecture of the CRISPR loci

(Makarova et al. 2011). The cas1 and cas2 genes, presumably involved in spacer integration (Brouns et al. 2008), are present in all (active) CRISPR/Cas systems and occur in every (sub) type (Makarova et al. 2011). The typical type 1 CRISPR/Cas system additionally includes a cas3 gene, which is composed of a helicase and a nuclease domain (Haft et al.

2005, Makarova et al. 2006) and is required for elimination of the invading DNA (Brouns et al. 2008, Sinkunas et al. 2011), as well as genes encoding a Cascade-like protein complex (Makarova et al. 2011). These complexes process the primary CRISPR transcript into small crRNAs (Brouns et al. 2008, Haurwitz et al. 2010, Jore et al. 2011). They contain several proteins belonging to the RAMP superfamily, which include the Cas5, Cas6, and Cas7 protein families, where a Cas6 variant most often is the enzyme in the complex that exhibits the RNA endonuclease activity (Makarova et al. 2011). In contrast, in the type II CRISPR/Cas system, crRNA maturation involves a trans-encoded small RNA called tracrRNA with sequence similarity to the repeat regions of the crRNA precursor, which directs the processing of the precursor by RNase III and Cas9 (Csn1 according to Haft et al. 2005) (Deltcheva et al. 2011). It is likely that Cas9 is also responsible for target

cleavage (Makarova et al. 2011). For the CRISPR/Cas type II system of Streptococcus thermophilus cleavage of phage and plasmid DNA within the proto-spacer has been shown in vivo (Garneau et al. 2010). Less is known about the type III CRISPR/Cas system; it contains polymerase and RAMP modules, and can be further divided into subtype III-A and subtype III-B (Makarova et al. 2011). Staphylococcus epidermidis is an example of a Type III-A system, and has been shown to target plasmid DNA in vivo (Marraffini and

Sontheimer 2008) while for the type III-B system of Pyrococcus furiosus in vitro evidence points in the direction of RNA molecules being the target (Hale et al. 2009).

CRISPR/Cas systems are thus widely spread in prokaryotic species and very diverse in their nature.

The model organism for this study, E. coli K12, has a type I CRISPR/Cas system. There

are 3 CRISPR loci within the E. coli K12 chromosome, CRISPR-I has 12 spacer/repeat

units, and a cluster of 8 cas genes is located immediately upstream (Figure 1). The

CRISPR-II and -III have 6 and 2 spacer/repeat units respectively and no associated cas

genes. No spacer acquisition or target interference with non-manipulated E. coli CRISPR

loci has been documented (Mojica and Diez-Villaseñor 2010). Low levels of Cascade are

limiting the CRISPR mediated defense in wild type E. coli (Pougach et al. 2010, Westra et

al. 2010). When challenging a population with phage, Pougach et al. (2010) noted a

limited CRISPR interference; λ-immunized strains formed smaller plaques than wild type

E. coli. High levels of immunity against phages can be obtained by overexpression of the

E. coli CRISPR/Cas system from recombinant plasmids (Brouns et al. 2008).

(6)

The aim of this project was to advance our understanding of CRISPR/Cas mediated immunity, in particular the interference stage using the type I E. coli CRISPR/Cas system as model. We modified the E. coli chromosome to create a CRISPR active E. coli strain.

Investigation include PAM requirement for CRISPR interference and visualizing the site- specific cleavage of target DNA.

Results

Creation of an active E. coli CRISPR/Cas system

Since the CRISPR/Cas system appears to be naturally inactive in E. coli K12 due to low levels of Cascade (Pougach et al. 2010, Westra et al. 2010), we inserted a kanamycin resistance cassette between cas3 and casA on the chromosome. The promoter of the Kan resistance gene is expected to read though the entire Cascade operon, cas1, cas2 and possibly into the CRISPR array, boosting expression levels (Figure 2A). Strains were transformed with a plasmid containing cas3 under a PLlacO-1 promoter (Figure 2C) or an arabinose inducible pBAD promoter (Figure 2C). To immunize the strain against phage lambda, an artificial CRISPR array targeting the template strand of 4 essential lambda genes J, O, R, E, was introduced in the strain replacing the wt spacers 1, 3, 5 and 7 and introducing restriction sites in spacers 2, 4 and 6 (EcoRI, BamHI and NsiI,

respectively) (Figure 2D). This artificial CRISPR array has previously been shown to convey immunity to phage lambda when expressed from a plasmid by Brouns and co- workers (2008).

Characterization of the modified E. coli CRISPR strain

To test if the modifications made were effective and the engineered strains containing spacers matching the lambda genome were insensitive to infection, plaque assays and growth curve analysis of lambda-infected cells were made.

The results were initially promising. The plaque assays showed an efficiency of plating (EOP) of 0 for λ-spacer containing cells compared to those still harboring the wt- spacers. For those strains harboring the plasmid with cas3 under the pBAD promoter immunity was inducible by addition of arabinose to the growth media, EOP 0 with arabinose and 0.87 without.

The growth curve experiments gave inconsistent results. Initially growth curves, in line with the plaque assays, showed CRISPR immunized cultures growing similar to an un- infected control while non-immune cultures lysed readily (figure 3A and B). When using pBAD-controlled cas3 this immunity was inducible by addition of arabinose to the growth media (figure 3B). However, these results were not repeatable and subsequent growth curves showed that cells provided with λ-spacers lysed upon infection with phage lambda (figure 3C).

The reversion of λ-immunized E. coli cells to sensitive cells can have several causes.

Mutation in either phage or spacers is unlikely since all four spacers or proto-spacers present would have to be mutated for complete loss of immunity to occur. The phage stock used was tested against the strains described by Brouns et al. (2008) from which the artificial lambda-CRISPR array was derived and these cells were still immune

(supplementary data) indicating that proto-spacers were still intact.

(7)

Figure 2; The active CRISPR system. A. A Kanamycin resistance cassette inserted between cas3 and casA on the E. coli chromosome. B. A kanamycin resistance cassette is inserted between cas3 and casA on the E. coli chromosome, and wt- CRISPR locus is replaced with an artificial λ-CRISPR array. C. Plasmid containing cas3 under the control of a PLlacO-1 promoter. D. Plasmid containing cas3 under the control of a pBAD promoter.

Different culture conditions such as numbers of phage used or the induction of PLlacO- 1 driven cas3 expression by addition of IPTG did not seem to affect immunity

substantially (supplementary data).

An upstream ATG/AAG motif seems important for interference

PAMs, or proto-spacer adjacent motifs, were first identified in S. thermophilus (Deveau et al. 2008, Horvath et al. 2008). They are, as the name implies, sequence motifs associated with the spacer precursor and might function as a recognition motif for spacer selection.

Phage can avoid the CRISPR/Cas system by mutating the PAM sequence (Deveau et al.

2008) and, in Sulfolobus (CRISPR type I-C), constructs carrying proto-spacers matching a spacer produced few transformants when flanked by their correct PAM motif while high transformation levels were observed with variations of the PAM motif (Gudbergsdottir et al. 2011), indicating the PAM has an important role in CRISPR interference in these strains. Using bioinformatic analysis, Mojica and co-workers identified the PAM for our model organism E. coli to be ATG or AAG immediately upstream of the spacer sequence in its transcriptional direction (Mojica et al. 2009). In other words, the PAM should be

CRISPR - array

L

!"!##$$$$!$!$$"!$!!!!"#"""$$!

!"!##$$$$!$!$$"!$!!!!"#"""$$!

!"!##$$$$!$!$$"!$!!!!"#"""$$!

!"!##$$$$!$!$$"!$!!!!"#"""$$!

!"!##$$$$!$!$$"!$!!!!"#"""$$!

!"!##$$$$!$!$$"!$!!!!"#"""$$!

!"!##$$$$!$!$$"!$!!!!"#"""$$!

!"!##$$$$!$!$$"!$!!!!"#"""$$!

!"!##$$$$!$!$$"!$!!!!"#"""$$!

!"!##$$$$!$!$$"!$!!!!"#"""$$!

!"!##$$$$!$!##"!$!!!!"#"""$$!

!"!##$$$$!$!$$"!$"!!!"#"""$$!

!#!##$$$$!$"#$"!$!!!!"#"""$$!

$###$!$"!"$!$!$!!$!"#"$!$#$"$!$"

$"!$$!""!$$"""!!#!"#!$$!""$"$!$#

!!$#$$$#!#$!!##!#""##!"#""#!##!"

###!!"#$!!!#$#!!""###$#!"!$!!#$!$

$!""#$!$!$"#"$$$#!$!$!#$!$$!$$#!$

#$"!$###"#"""#$$!!"!"#"$!!"""$#"

!"$#$"$$$$!"""!"!"##!$$"!$$"!$##

$#!$#!!"!$#!!$#!$""!!$""!$$!$$$"

!!!!!$!$"#!"$$!#"""$"##"#$$$$$!!

!!"!##$"!"$"#"!!#!!""#!"#!!"$#"$

$$$!!#"!$$"!!###!$""$!$$#!""$$!"

!$""$!"$!!#!"!"###$"$!$$#!"$!$#!

Repeats Spacers

1 2 3 4 5 6 7 8 9 10 11 12

cas3 PkanR kanR casA

(cse1) casB (cse2) casC

(cas7) casD (cas5) casE

(cas6e)cas1 cas2

cas3 casA

(cse1) casB (cse2) casC

(cas7) casD (cas5) casE

(cas6e)cas1 cas2 CRISPR - array

L

!"!##$$$$!$!$$"!$!!!!"#"""$$!

!"!##$$$$!$!$$"!$!!!!"#"""$$!

!"!##$$$$!$!$$"!$!!!!"#"""$$!

!"!##$$$$!$!$$"!$!!!!"#"""$$!

!"!##$$$$!$!$$"!$!!!!"#"""$$!

!"!##$$$$!$!$$"!$!!!!"#"""$$!

!"!##$$$$!$!$$"!$!!!!"#"""$$!

$#!"!#!#!"#$!"#!$$"#$"!$!""!!!$$

$"!$$!""!$$"""!""##$!$$!""$"$!$#

$""!$""$"!!$"!!$!#!"$"!$$"!$"""$

###!!"#$!!!#$#!!"#$$#$#!"!$!!#$!

#!!!"#!$$#"$$!$""!$"!$##!!$$#!""

#$"!$###"#"""#"#!$"#"#"$!!"""$#"

#!"$""!#$$"$!#"#!"$$$!"$$!"$!"#"

Repeats Spacers

λ-J EcoRI

λ-O BamHI

λ-R NsiI λ-E PkanR kanR

A

B

C D

(8)

Figure 3; Growth curves after phage-λ infection and transformation efficiency of CRISPR activated E. coli W3110. A.

Growth curve showing normalized OD600 values of E. coli W3110 KanR /pNH42 (Control) and E. coli W3110 KanR λ-CRISPR array /pNH42 (Active) infected with λvir at different virus concentrations and including a non-infected control.

The E. coli harboring the λ-CRISPR spacers are growing as the un-infected control at multiplicity of infection (MOI) 0.1 and 1 (red and green line respectively) B. Growth curve showing the average normalized OD600 values of three independent experiments for E. coli W3110 KanR /pNH43 (Control) and E. coli W3110 KanR λ-CRISPR array /pNH43 (Active) infected with λvir (indicated by a + in the second row of the legend, Turquoise, Orange, light-blue and pink lines) and un-infected controls (indicated by a – in the second row of the legend, dark-blue, red, green, purple lines) in the presence of arabinose (+ in the first row of the legend, red, purple, orange, pink lines) or absence of arabinose (- in the first row of the legend, dark-blue, green, turquoise, light-blue lines). Immunity to phage lambda is inducible by addition of arabinose; compare active strain with arabinose (pink line) to active strain without arabinose (light blue line). C. Growth curve showing the average normalized OD600 values of three independent experiments for E. coli W3110 KanR /pNH42 (Control) and E.

coli W3110 KanR λ-CRISPR array /pNH42 (Active) infected with λvir with virus concentration to MOI 1. The E. coli W3110 KanR λ-CRISPR array /pNH42 active strain (red line) that was insensitive to lambda infection before (compare green line from figure A) now behaves as sensitive control strain (purple line). D. Transformation of target plasmid containing wt-spacer 3 flanked by the trinucleotide sequences indicated in the figure into E. coli W3110 KanR /pNH34 (Active, red bars) and W3110 KanR λ-CRISPR array /pNH34 (Control, blue bars) cultured in presence of arabinose. The proto spacer is indicated by ---, so CTT --- CTT means there is a CTT motif present both upstream and downstream of the proto- spacer. AAC is placed between parenthesis to signify is has not been proposed as a PAM motif. Transformation efficiency is expressed as percentage relative to the transformation efficiency of a control plasmid (100%, not shown in graph) containing a random sequence instead of the E. coli wt-spacer 3.

found on the side of the spacer that is oriented towards the leader sequence. In the over- expression system used by Brouns and co-workers immunity against phage lambda was observed even though no ATG or AAG PAM was associated with the proto-spacers of their artificial λ-spacers (Brouns et al. 2008). While this could be due to overexpression, Pougach et al. (2010) detected partial interference in absence of a PAM.

To establish a PAM requirement for CRISPR interference in E. coli in vivo,

transformation efficiencies of various target plasmids into our modified CRISPR active E. coli strain were investigated. Target plasmids were constructed to have a proto-spacer homologous to the wild type CRISPR spacer 3 in various contexts: upstream ATG, CTT, or AAC (not a PAM motif), downstream AAG, CAT, and upstream as well as

A B

C D

(9)

downstream AAG, CTT motifs (see figure), and equipped with a chloramphenicol resistance gene for selection. Transformation assays were made using CRISPR active W3110 E. coli strains containing an arabinose inducible cas3 expression vector. As negative control we used the same CRISPR active W3110 E. coli strain but with the wt- CRISPR locus exchanged for artificial λ-CRISPR spacers and the wt-spacer 3 absent.

Transformation experiments were done with cas3 induced and repressed cells, but only data for the induced system is shown.

Transformation efficiency of the target plasmid containing proto-spacer 3 is expressed as a percentage of the transformation efficiency from a random sequence control in the exact same genetic context on the plasmid. If the plasmid is eliminated by the

CRISPR/Cas system we would expect low transformation efficiency, and this is exactly what we see for those constructs that have the predicted AAG or ATG motif present upstream of the proto-spacer, with a transformation efficiency of 0% in cells having a complementary spacer. In cells lacking this spacer, transformation efficiency is more than 100% indicating that the effect is CRISPR dependent (figure). Those plasmids having the PAM-like AAC upstream of the spacer, targeting was less efficient but still substantial; a transformation efficiency of 2, 6 and 15% for spacers with a downstream GAG, CAT and AAG respectively (figure). Plasmids carrying proto-spacer 3 flanked by CTT upstream are transformed readily into the CRISPR activated E. coli despite the presence of the matching spacer and thus seem to avoid targeting by the CRISPR/Cas system;

transformation efficiency >100% (figure). From these results it can be concluded that an upstream ATG, AAG or a closely related motif is necessary for interference by the type I-E (or E. coli) CRISPR/Cas system.

Discussion

In this initial in vivo investigation of the targeting phase of the CRISPR/Cas type 1-E (previously known as Ecoli subtype, Haft et al. 2005) we have shown that the upstream ATG or AAG PAM motif previously only identified by bioinfomatics (Mojica et al. 2009) or a similar AAC motif is needed for CRISPR interference against plasmids. This is in contrast to earlier results from Brouns et al. (2008) where they reported high levels of protection against phage lambda using a Cas over-expression system and artificial λ- spacers where a PAM did not flank the proto-spacers. Investigation shows that the proto-spacers of the artificial λ-spacers Jt, Ot, Rt (targeting the template strand of the J O R genes on the lambda chromosome) used by Brouns et al. (2008) are flanked by TTG, TAA, and TGG respectively. The inactive Et spacer (Brouns et al. 2008) has GCG immediately upstream of the proto-spacer. These results agree well with the finding of Marraffini and Sontheimer (2010) that at least 2 consecutive mismatches between the repeat on the crRNA and the sequence immediately upstream of the proto-spacer on the target DNA at positions -4 to -2 are required for elimination. The ATG and AAG PAMs found by Mojica et al. (2009) and the Jt TTG and Rt TGG motifs follow this rule; they have mismatches at positions -2 and -3 and a match at -1 with the E. coli repeat sequence CCG. The AAC motif we found as well as the GTA flank of spacer Ot from Brouns et al. (2008) have three consecutive mismatches from positions -1 till -3. While base pairing at the -1 position seems to be preferred, it does not seem to be essential to targeting.

A possible explanation of the discrepancy between the PAMs found by Mojica et al.

(2009) and the PAM-like motifs found in this study is a possible difference in specificity between the acquisition of new spacers and the targeting of foreign DNA. That PAMs have a role in spacer acquisition has previously been speculated by others (Deveau et al.

2008, Horvath et al. 2008, Mojica et al. 2009). The process of spacer acquisition might not

(10)

tolerate the mismatch at the -1 position and thus such motifs do not show up among naturally occurring PAMs. The specificity of the targeting complex might be slightly different, tolerating PAM-like sequences as well as the more stringent PAMs. The motifs tested in this study were only a small subset of the possible combinations; a more

thorough investigation might give a better understanding of the sequence requirements of PAMs for CRISPR interference.

We created an E. coli strain with an activated CRISPR/Cas system to be able to show cleavage of target phage and plasmid DNA using Southern blot and primer extension analysis. Several problems were encountered in our attempts. The first and foremost problem were the inconsistent results obtained when doing growth-curve analysis of the engineered strains; while initially strains appeared insensitive to infection with λ

vir

these results were not reproducible. The cause of the immunity loss could not be identified, and it seems that a careful reevaluation of the system design is needed. Secondly a problem occurs when using vectors under selective pressure when investigating targeting of plasmid DNA. As Gudbergsdottir et al. (2011) have shown in Sulfulobus when

introducing a proto-spacer-carrying vector maintained under selection, the majority of transformants will die due to uracil deficiency because in their case selection was achieved using a mutant Sulfulobus strain unable to synthesize uracil and compensating this mutation by supplementing the necessary genes on the plasmid. This is reminiscent of what we see in E. coli for our transformation efficiency assays where selection is based on antibiotic resistance genes carried on the target plasmid. Some transformants,

however, will survive despite selection, mostly due to deletions in the CRISPR locus that include the matching spacer, or as a result of complete deletion of the CRISPR/Cas module (Gudbergsdottir et al. 2011). This means that we may be selecting for those CRISPR-inactive mutants when culturing the cells with target plasmid in antibiotic containing medium. To overcome this problem, the CRISPR/Cas system needs to be tightly regulated, so that we can culture without selecting for CRISPR mutants and switch on the system once ready for sampling. In the current system, cas3 is mainly expressed from plasmids and when using the pBAD construct, expression is repressed by addition of glucose and induced by addition of arabinose. However the chromosomal copy of cas3 is still intact and expressed at a low rate (unpublished real-time pcr data), which might be enough for selection. A chromosomal deletion of cas3 would solve this issue. And, last but not least, cleaved target DNA has to be isolated from the cells before further analysis, but exonucleases present in the E. coli cells might degrade the linearized target DNA. To address this, we have made recD knockout variants of the engineered strains.

Methods

Bacterial strains and culture conditions

The Escherichia coli K12 W3110 strain was used throughout this study. Cells were cultured in Luria-Bertani (LB) medium, whenever required suitable antibiotics were added to the medium at the following final concentrations: ampicillin 100 µg/ml, kanamycin 50 µg/ml, chloramphenicol 30 µg/ml, tetracycline 12 µg/ml. When preparing cells for plaque assays or growth-curve analysis the LB medium was complemented with 0.2%

maltose and 10mM MgSO

4

for efficient lambda infection.

(11)

Construction of a CRISPR active E. coli strain

To immunize E. coli against λ-phage, a synthetic CRISPR array of four spacers targeting the template strand of the lambda J, O, R, and E genes described by Brouns et al. (2008) was inserted on the chromosome of E. coli W3110 of which the wild type CRISPR locus I had previously been knocked out, using the lambda red recombineering system (first described by Datsenko and Wanner 2000). A kanamycin resistance cassette (described Datsenko and Wanner, 2000) was inserted in the transcriptional direction between cas3 and casA on the chromosome of the constructed E.coli W3110 λ-CRISPR and a wild type E. coli W3110 strain using lambda red recombineering with primers R12 and AR123, generating the two prime constructs for this research; E.coli W3110 KanR, λ-CRISPR (N279) and E.coli W3110 KanR (N277). These constructs were subsequently made competent by using the CaCl

2

method, and transformed with one of the vectors carrying the E. coli cas3 gene that are described below.

pNH42 carried cas3 under the control of a PLlacO-1 promoter. To create this vector the cas3 gene was amplified using the primers NH309 and NH303. The NH303 incorporates an XbaI site downstream of cas3. The NH309 primer is phosphorylated to enable blunt end ligation. The pZE12 linear backbone was generated by PCR using the primers pLacOC-P and pLacOB, and the obtained product was treated with DpnI to degrade the template plasmid. After digestion of insert and vector with XbaI the resulting parts could be ligated.

The arabinose inducible cas3 expression vector pNH43 was generated by cloning the cas3 gene fragment, obtained by XbaI cleavage of pNH8 (pCF430 with E. coli cas3 using primers NH198 and 199) subsequent blunting of the overhangs before digestion with EcoRI and excising the approximately 3kb fragment from agarose gel, into the 13.2 kb EcoRI, PvuII digestion product of the pCU01 vector which is a pBAD-TOPO vector derivate (Unoson and Wagner, 2008).

Another pBAD driven cas3 expression vector, pNH34, was constructed by cloning a cas3 amplicon from the E. coli chromosome generated with primers NH198 and NH275 into pCF430 using the XbaI and KpnI restriction enzymes.

Phage studies

Host sensitivity was tested using a virulent lambda phage variant (λ

vir

) by plaque assays, and growth-curve analysis (described below). Propagation of λ

vir

was done as described by Patterson and Dean 1987 but host used for propagation was E. coli K12 MG1655, and the culture was monitored by taking OD measurements every 30 minutes, phage were harvested immediately upon lysis of the culture, phage titer was determined by plaque assay. The host strains for infection were the engineered E. coli K12 W3110 strains described above (E. coli K12 W3110 AR122/123 KanR and E. coli K12 W3110

AR122/123 KanR, λ-CRISPR template with the pNH42 or pNH43 plasmids. See table S1).

When preparing plating cells for the plaque assays of the strains harboring the pNH43 plasmid with pBAD controlled cas3 expression an overnight culture was diluted 1 in 100 in 50 ml LB supplemented with 0.2% maltose and 10mM MgSO

4

and grown to OD600

~ 0.3, the culture was then centrifuged at room temperature (RT) 3000 rpm for 4

minutes, resuspended in 50 ml pre-warmed LB containing 0.2% maltose, 10mM MgSO

4

(un-induced) or the same but additionally containing 0.2% arabinose (induced), after 30

(12)

minutes of induction complete cultures were centrifuged (RT, 3000 rpm, 4 minutes) and pellet resuspended in 5ml 10mM MgSO

4

. Plates and top agar used contained 0.2%

maltose, 10mM MgSO

4

and for induced cells additionally 0.2% arabinose.

Calculation of the efficiency of plating (EOP) of the immunized λ-CRISPR containing strains compared to those strains still harboring the wt-CRISPR array was done as follows; the plates were scored, and by dividing the number of plaques through the dilution factor, the plaque forming units (pfu) per ml is obtained, the pfu/ml of the strains with λ-CRISPR template were than divided by the pfu/ml of the complementing strain with the wt-CRISPR template giving the EOP of the active λ-CRISPR strains.

For growth curve analysis an overnight culture was diluted 1:100 in LB containing 0.2%

maltose and 10mM MgSO

4

and grown till OD600 ~ 0.3 before λ

vir

was added. OD600 measurements were made up till 4 or 5 hours after addition of phage at 60 minutes intervals. When possible curves shown are the average of three independent experiments, for this the absolute OD600 measurements were normalized by expressing the values obtained relative to the start OD just before addition of λ

vir

(Divide obtained value with start value, which is then set to 1). Whenever the multiplicity of infection (MOI) is indicated the amount of phage stock to be added was calculated by dividing the total amount of cells in the culture (to convert OD600 values to cells/ml a calibration curve was made using E. coli K12 W3110, a OD600 of 0.1 was found to approximate 10

7

cells) through the pfu/ml value obtained for the phage stock (by plaque assay).

Construction of plasmids for transformation studies

The target plasmid is a low-copy vector that carries a pSC101 origin of replication (3–4 plasmid copies/cell), a chloramphenicol resistance marker, and the target sequence complementary to the wild type CRISPR spacer 3 as a transcriptional fusion upstream of a gfp gene.

Transformation efficiency studies

For transformation efficiency studies E. coli K12 W3110 AR122/123 KanR and E. coli K12 W3110 (active) AR122/123 KanR, λ-CRISPR template (control) with the arabinose inducible cas3 expression vector pNH34 (see construction of CRISPR active E. coli strain and table S). To see if CRISPR/Cas functioning was inducible by addition of arabinose to the media cells were grown with either arabinose (induced) or glucose (repressed) present.

Cells were made CaCl

2

competent by inoculating 60 ml LB containing either 0.2%

glucose (repressed) or 0.2% arabinose (induced), with 600µl of an overnight culture grown in presence of glucose. Cultures were grown for 90 minutes till OD600 ~ 0.3 and each distributed in six 15ml falcon tubes, cells were harvested by centrifugation

(6000rpm, 2 min at 4°C), and resuspended in 2 ml ice-cold 0.1 M CaCl

2

. After 30 minutes of incubation on ice the cell suspension was centrifuged (13000 rpm, 1 min at RT) and pelleted cells carefully resuspended in 600µl ice-cold 0.1 M CaCl

2

. After 2 more hours of incubation on ice glycerol was added to a final concentration of 10%, aliquots of 50µl were made and stored at -80°C.

For transformation the CaCl

2

competent cells were thawed on ice, and incubated with

~20-85 ng plasmid DNA for 20 minutes before being subjected to heat-shock (2 min,

42°C), after which 1 ml LB (with either 0.2% glucose for repressed or 0.2% arabinose for

(13)

induced cells) at RT was added. After shaking at 37°C for 1 hour 100 µl was plated on selective media. Plates were scored the following day and transformation efficiency calculated using the following formula;

# Colonies on plate/ ng of DNA plated (~2-8,5 ng)

Transformation efficiency of each target plasmid was expressed as a percentage relative to the transformation efficiency of a random sequence control having the exact same genetic context (divide the transformation efficiency of target though that of the control).

Acknowledgements

First and foremost I would like to thank prof. Gerhart Wagner for giving me the

opportunity to work in his group and for his good advice and enthusiasm during the last year. I would like to thank my supervisor Nadja Heidrich for her help and support. I thank Amanda Raine, Magnus Lundgren, Erik Holmqvist and Sriram Kurella for helpful discussions. And last but not least Erik Gullberg for long discussions over coffee and for the help with the illustrations.

References

Al-Attar, S., Westra, E. R., van der Oost, J., & Brouns, S. J. (2011). Clustered regularly interspaced short palindromic repeats (CRISPRs): the hallmark of an ingenious antiviral defense mechanism in prokaryotes. Biological Chemistry , 392 (4), 277-289.

Barrangou, R., Fremaux, C., Deveau, H., Richards, M., Boyaval, P., Moineau, S., et al.

(2007). CRISPR provides acquired resistance against viruses in prokaryotes. Science , 315, 1709-1712.

Bolotin, A., Quinquis, B., Sorokin, A., & Ehrlich, S. D. (2005). Clustered regularly interspaced short palindrome repeats (CRISPRs) have spacers of extrachromosomal origin. Microbiology , 151, 2551-2561.

Brouns, S. J., Jore, M. M., Lundgren, M., Westra, E. R., Slijkhuis, R. J., Snijders, A. P., et al. (2008). Small CRISPR RNAs Guide Antiviral Defense in Prokaryotes. Science , 321 (5891), 960-964.

Datsenko, K. A., & Wanner, B. L. (2000). One-step inactivation of chromosomal genes in Escherichia coli K-12 using PCR products. PNAS , 97 (12), 6640-6645.

Deltcheva, E., Chylinski, K., Sharma, C. M., Gonzales, K., Chao, Y., Pirzada, Z. A., et al.

(2011). CRISPR RNA maturation by trans-encoded small RNA and host factor RNase III. Nature , 471 (7340), 602-607.

Deveau, H., Barrangou, R., Garneau, J., Labonté, J., Fremaux, C., Boyaval, P., et al.

(2008). Phage Response to CRISPR-Encoded Resistance in Streptococcus thermophilus.

Journal of Bacteriology , 190 (4), 1390-1400.

Deveau, H., Garneau, J. E., & Moineau, S. (2010). CRISPR/cas system and its role in phage-bacteria interactions. Annual Review of Microbiology , 64, 475-493.

Garneau, J. E., Dupuis, M.-E., Villion, M., Romero, D. A., Barrangou, R., Boyaval, P., et

al. (2010). The CRISPR/Cas bacterial immune system cleaves bacteriophage and plasmid

DNA. Nature , 468 (7320), 67-71.

(14)

Grissa, I., Vergnaud, G., & Pourcel, C. (2007). The CRISPRdb database and tools to display CRISPRs and to generate dictionaries of spacers and repeats. BMC Bioinformatics , 8 (172), doi: 10.1186/1471-2105-8-172.

Groenen, P. M., Bunschoten, A. E., van Soolingen, D., & van Embden, J. D. (1993).

Nature of DNA polymorphism in the direct repeat cluster of Mycobacterium

tuberculosis; application for strain differentiation by a novel typing method. Molecular Microbiology , 10 (5), 1057-1065.

Gudbergsdottir, S., Deng, L., Chen, Z., Jensen, J. V., Jensen, L. R., She, Q., et al. (2011).

Dynamic properties of the sulfolobus CRISPR/Cas and CRISPR/Cmr systems when challenged with vector-borne viral and plasmid genes and protospacers. Molecular Microbiology , 79 (1), 35-49.

Haft, D. H., Selengut, J., Mongodin, E., & Nelson, K. E. (2005). A guild of 45 CRISPR- associated (Cas) protein families and multiple CRISPR/Cas subtypes exist in prokaryotic genomes. PLoS Computational Biology , 1 (6), 0474-0483.

Hale, C. R., Zhao, P., Olson, S., Duff, M. O., Graveley, B. R., Wells, L., et al. (2009).

RNA-guided RNA cleavage by a CRISPR RNA-Cas protein complex. Cell , 139 (5), 945- 956.

Horvath, P., & Barrangou, R. (2010). CRISPR/Cas, the immune system of bacteria and archaea. Science , 327 (5962), 167-170.

Horvath, P., Romero, D. A., Coute-Monvoisin, A.-C., Richards, M., Deveau, H., Moineau, S., et al. (2008). Diversity, Activity, and Evolution of CRISPR Loci in Streptococcus thermophilus. Journal of Bactriology , 190 (4), 1401-1412.

Ishino, Y., Shinagawa, H., Makino, K., Amemura, M., & Nakata, A. (1987). Nucleotide sequence of the iap gene, responsible for alkaline phosphatase isozyme conversion in Escherichia coli, and identification of the gene product. Journal of Bacteriology , 169 (12), 5429-5433.

Jansen, R., van Embden, J. D., Gaastra, W., & Schouls, L. M. (2002). Identification of genes that are associated with DNA repeats in prokaryotes. Molecular Microbiology , 43 (6), 1565-1575.

Jore, M. M., Lundgren, M., van Duijn, E., Bultema, J. B., Westra, E. R., Waghmare, S. P., et al. (2011). Structural basis for CRISPR RNA-guided DNA recognition by Cascade.

Nature structural & molecular biology , 15 (5), 529-536.

Kunin, V., Sorek, R., & Hugenholtz, P. (2007). Evolutionary conservation of sequence and secondary structures in CRISPR repeats. Genome Biology , 8 (4), doi: 10.1186/gb-2007- 8-4-r61.

Makarova, K. S., Grishin, N. V., Shabalina, S. A., Wolf, Y. I., & Koonin, E. V. (2006). A putative RNA-interference-based immune system in prokaryotes: computational analysis of the predicted enzymatic machinery, functional analogies with eukaryotic RNAi, and hypothetical mechanisms of action. Biology Direct , 1 (7), doi:10.1186/1745-6150-1-7.

Makarova, K. S., Haft, D. H., Barrangou, R., Brouns, S. J., Charpentier, E., Horvath, P.,

et al. (2011). Evolution and classification of the CRISPR-Cas systems. Nature reviews

Microbiology , 9 (6), 467-477.

(15)

Marraffini, L. A., & Sontheimer, E. J. (2008). CRISPR interference limits horizontal gene transfer in Staphylococci by targeting DNA. Science , 322, 1843-1845.

Marraffini, L. A., & Sontheimer, E. J. (2010). CRISPR interference: RNA-directed adaptive immunity in bacteria and archaea. Nature Reviews Genetics , 11, 181-190.

Mojica, F. J., & Diez-Villasenor, C. (2010). The on-off switch of CRISPR immunity against phages in Escherichia coli. Molecular Microbiology , 77 (6), 1341-1345.

Mojica, F. J., Diez-Villasenor, C., Garcia-Martinez, J., & Almendros, C. (2009). Short motif sequences determine the targets of the prokaryotic CRISPR defence system.

Microbiology , 155 (3), 733-740.

Mojica, F. J., Diez-Villasenor, C., Garcia-Martinez, J., & Soria, E. (2005). Intervening sequences of regularly spaced prokaryotic repeats derive from foreign genetic elements.

Journal of Molecular Evolution , 60 (2), 174-182.

Mojica, F. J., Ferrer, C., Juez, G., & Rodriguez-Valera, F. (1995). 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 , 17 (1), 85-93.

Patterson, T. A., & Dean, M. (1987). Preparation of high titer lambda phage lysates.

Nucleic acids research , 15 (15), 6298.

Pougach, K., Semenova, E., Bogdanova, E., Datsenko, K. A., Djordjevic, M., Wanner, B.

L., et al. (2010). Transcription, processing and function of CRISPR cassettes in Escherichia coli. Molecular Microbiology , 77 (6), 1367-1379.

Pourcel, C., Salvignol, G., & Vergnaud, G. (2005). CRISPR elements in Yersinia pestis acquire new repeats by preferential uptake of bacteriophage DNA, and provide additional tools for evolutionary studies. Microbiology , 151 (3), 653-663.

Pul, U., Wurm, R., Arslan, Z., Geißen, R., Hofmann, N., & Wagner, R. (2010).

Identification and characterization of E. coli CRISPR-cas promoters and their silencing by H-NS. Molecular Microbiology , 75 (6), 1495-1512.

She, Q., Singh, R. K., Confalonieri, F., Zivanovic, Y., Allard, C., Awayez, M. J., et al.

(2001). The complete genome of the crenarchaeon Sulfolobus solfataricus P2. PNAS , 98 (14), 7835-7840.

Sinkunas, T., Gasiunas, G., Fremaux, C., Barrangou, R., Horvath, P., & Siksnys, V.

(2011). Cas3 is a single-stranded DNA nuclease and ATP-dependent helicase in the CRISPR/Cas immune system. The EMBO Journal , 30 (7), 1335-1342.

Stern, A., Keren, L., Wurtzel, O., Amitai, G., & Sorek, R. (2010). Self-targeting by CRISPR: gene regulation of autoimmunity? Trends in Genetics , 26 (8), 335-339.

Unoson, C., & Wagner, E. G. H. (2008). A small SOS-induced toxin is targeted against the inner membrane in Escherichia coli. Molecular Microbiology , 70 (1), 258-270.

van der Oost, J., Jore, M. M., Westra, E. R., Lundgren, M., & Brouns, S. J. (2009).

CRISPR-based adaptive and heritable immunity in prokaryotes. Trends in Biochemical

Sciences , 34 (8), 401-407.

(16)

Westra, E. R., Pul, U., Heidrich, N., Jore, M. M., Lundgren, M., Stratmann, T., et al.

(2010). H-NS-mediated repression of CRISPR-based immunity in Escherichia coli K12

can be relieved by the transcription activator LeuO. Molecular Microbiology , 77 (6), 1380-

1393.

(17)

Glossary  

CRISPR/Cas system: A adaptive inheritable prokaryotic defense system against foreign genetic elements, consisting of one or possibly several CRISPR arrays flanked by a A/T rich leader sequence and a variable set of cas genes.

CRISPR array: A series of direct palindromic repeats interspaced by regularly sized variable spacer sequences usually derived from extrachromosomal elements. CRISPR stands for Clustered Regularly Interspaced Short Palindromic Repeats.

Repeat: A palindromic sequence repeated within a CRISPR array.

Spacer: Variable sequence in a CRISPR array, flanked by repeats. Spacers usually derive from extrachromosomal elements such as phages or plasmids.

Proto-spacer: the sequence in the target DNA that is homologous to the spacer.

PAM: Proto-spacer Adjacent Motif, a di- or tri- nucleotide motif found immediately next to the proto-spacer in the target DNA.

Leader sequence: A/T rich sequence directly upstream of the CRISPR array.

crRNAs: small RNA derived from the primary CRISPR transcript that guides the defense apparatus to the target nucleic acid.

Cascade: CRISPR-associated complex for antiviral defense. Protein complex containing

the CasABCDE that processed the primary CRISPR transcript into crRNAs in E. coli

References

Related documents

Phylogenetic group distributions, virulence factors and antimicrobial resistance properties of uropathogenic Escherichia coli strains isolated from patients with

coli reference collection (ECOR) (Ochman and Selander, 1984) but also two clinical ESBL-harboring E. From our study, we suspect that the process of phage infection and

(1999) Escherichia coli mutants lacking all possible combinations of eight penicillin binding proteins: viability, characteristics, and implications for peptidoglycan

Oligonucleotides were designed according to the CRISPR array used in Cascade purification (Figure 11; Weiss A, 2009. CRISPR- associated protein complex from Escherichia

Partial deletions of the region upstream of qepA revealed that the insertion sequence IS26 is not required for the resistance phenotype, but deletion of the integrase int1 and

Syftet med denna studie var att ta ställning till om EHEC-PCR ska införas som rutinmetod eller användas parallellt med odlingsmetoden i utvärderingssyfte på de prover som kommer in

Therefore, this study investigated how experimental and observational data can be used in mechanistic and statistical models to improve predictions of bacterial transport

coli K-12 derivatives, a small subpopulation of bacterial cells (ca 1-2%) apparently expressed a high enough level of surface-exposed ClyA to be visible by immunofluores-