You are on page 1of 9

REVIEW ARTICLE

PUBLISHED: 5 JUNE 2017 | VOLUME: 2 | ARTICLE NUMBER: 17092

A decade of discovery: CRISPR functions


and applications
Rodolphe Barrangou1* and Philippe Horvath2
This year marks the tenth anniversary of the identification of the biological function of CRISPR–Cas as adaptive immune systems
in bacteria. In just a decade, the characterization of CRISPR–Cas systems has established a novel means of adaptive immunity
in bacteria and archaea and deepened our understanding of the interplay between prokaryotes and their environment, and
CRISPR-based molecular machines have been repurposed to enable a genome editing revolution. Here, we look back on the
historical milestones that have paved the way for the discovery of CRISPR and its function, and discuss the related technologi-
cal applications that have emerged, with a focus on microbiology. Lastly, we provide a perspective on the impacts the field has
had on science and beyond.

T
he nearly ubiquitous presence of diverse microorganisms and archaea. Indeed, years before the CRISPR acronym was actu-
throughout the planet is awe-inspiring, yet somewhat per- ally coined, a number of microbial geneticists came across these
plexing given the widespread occurrence of bacteriophages peculiar loci in the genomes of microorganisms, primarily focus-
(or phages) and viruses that infect them. Consequently, the constant ing on model organisms and pathogenic bacteria. Originally,
arms race between prokaryotes and viruses has yielded a powerful these series of interspaced and partially palindromic DNA repeats
and nimble arsenal of defence mechanisms, encompassing restric- were observed in an intergenic region upstream of the iap gene
tion modification, toxin–antitoxin modules, abortive infection and in Escherichia coli 5,6. Similar arrays of repeated DNA were also
the recently characterized CRISPR–Cas system1. The latter provides observed in human pathogens, such as Mycobacterium tuberculosis 7
adaptive immunity in bacteria and archaea (Box 1), and hinges on and Streptococcus pyogenes 8, as well as genomes of two Haloferax
diverse and versatile molecular machines that have popularized species9 and a filamentous cyanobacterium10. In these early days,
CRISPR not only as a compelling means to fend off viruses, but also several acronyms such as DVR (direct variable repeats), TREP (tan-
as a democratized technology broadly used in molecular biology. dem repeats), LTRR (long tandemly repeated repetitive sequences),
Here, we examine the various eras of CRISPR history, and highlight SRSR (short regularly spaced repeats), LCTR (large clusters of tan-
discoveries that have advanced the field. First, we discuss a series dem repeats), and later SPIDR (spacer interspersed direct repeats)
of milestone discoveries that assembled the CRISPR puzzle, and were used by various authors to refer to loci containing short and
basic findings that unravelled the molecular mechanisms provid- partially palindromic DNA tandem repeats interspaced by seem-
ing adaptive immunity in bacteria. Next, we provide an overview ingly random sequences. The advent of genome sequencing tech-
of the technical developments that have advanced the CRISPR field nologies in the 1990s was a tipping point providing a large set of
and generated the toolbox that enables a diverse set of historical and microbial genomes that carried these peculiar sequences, although
contemporary applications. Lastly, we discuss how these seminal their origin and function remained unknown.
discoveries and versatile technologies have impacted and are likely
to continue to shape the field of microbiology. Middle ages. In the early 2000s, with increasing availability of
genome sequences for numerous and diverse bacteria and archaea,
The humble beginnings and meteoric rise of CRISPR two groups consistently and repeatedly observed the occurrence
The remarkable number of studies dedicated to clustered regularly of CRISPR arrays2,3,11, leading to the creation of the consensus
interspaced short palindromic repeats (CRISPR)2 and CRISPR- acronym ‘CRISPR’ in 2002 (ref. 3) (Fig.  1). Though the biological
associated sequences (Cas)3 published in recent years illustrates the function of those widespread genetic features remained enigmatic
speed at which the field has evolved, and the scale at which CRISPR- for nearly 20 years after their initial discovery, several hypotheses
based technologies and applications have impacted the fields of were originally provided, ranging from replicon partitioning 9,11 to
microbiology, biology and genetics4. Yet CRISPR remained surpris- DNA repair 12.
ingly mysterious and underappreciated for decades; indeed, in the The first clue regarding the biological function of CRISPR was
beginning, CRISPR was a genomic peculiarity that was mostly an unearthed in 2005, when three distinct research groups nearly
assembly inconvenience. Here, we discuss the evolution of the per- concurrently made the observation that the seemingly random
ception of CRISPR and its significance by comparing the history of sequences separating identical CRISPR repeats actually showed
CRISPR to familiar historical periods in human history, leading to homology to invasive DNA sequences, such as viruses and plas-
unexpected, yet compelling, modern times (Fig. 1). mids13–15 (Fig.  1). It was then postulated that CRISPR may con-
stitute a putative defence system in prokaryotes13,16 that might
Antiquity. It may come as a surprise to many that the CRISPR lit- function analogously to the eukaryotic RNA interference (RNAi)
erature dates back to the late 1980s, when several groups repeat- system, even though there are no homologues of RNAi machinery
edly observed arrays of DNA repeats in the genomes of bacteria components among the Cas proteins16.

Department of Food, Bioprocessing, and Nutrition Sciences, North Carolina State University, 400 Dan Allen Drive, Campus Box 7624, Raleigh,
1

North Carolina 27695-7624, USA. 2DuPont Nutrition and Health, BP10, 86220 Dangé-Saint-Romain, France. *e-mail: rbarran@ncsu.edu

NATURE MICROBIOLOGY 2, 17092 (2017) | DOI: 10.1038/nmicrobiol.2017.92 | www.nature.com/naturemicrobiology 1


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEW ARTICLE NATURE MICROBIOLOGY

Box 1 | CRISPR mechanism of action.

There are two classes of CRISPR–Cas systems based on the processed into small, mature crRNAs that each contain a target-
nature of the effector nuclease that drives targeting: either a ing portion derived from a unique CRISPR spacer. Thirdly, dur-
multi-protein complex for class 1, or a single protein for class ing interference, crRNAs guide the Cas effector nuclease (yellow)
2. In either case, cas operons are adjacent to the CRISPR array, for sequence-specific targeting and cleavage of complementary
which is transcribed from the leader (L) promoter sequence. nucleic acid. Each type is defined by a dedicated Cas effector
Typically, cas1 and cas2 genes are present in most types and nuclease (see inset panel). In class 1, type I systems, the Cas3 exo-
subtypes across the two classes. For class 1 systems, a canoni- nuclease nicks and then chews the target DNA strand; in type III
cal type I CRISPR–Cas system is shown, with the signature Cas3 systems, the Cas10 nuclease cleaves mRNAs in a ruler–anchor
exonuclease and Cascade. For class 2 systems, a canonical type II mechanism manner, which is coupled in some systems to target
CRISPR–Cas system is shown, with the signature Cas9 endonu- DNA degradation; targeting remains uncharacterized in type IV
clease. First, during acquisition, a short viral DNA sequence is systems. In class 2 systems, type II systems use the Cas9 endo-
incorporated as a novel spacer (blue rectangle) into the CRISPR nuclease to generate two nicks that yield a dsDNA break; type V
array, where each repeat is shown as a black diamond. This is systems use Cas12 to generate two offset nicks; type VI systems
a Cas1–Cas2-dependent process, and novel spacers are inte- use Cas13 to generate a cut in the target mRNA and then pro-
grated in a polarized manner at the leader-end of the CRISPR cesses it non-specifically to yield collateral damage. For type I, II,
array. Secondly, during expression, the CRISPR array is tran- V and VI, interference relies on PAM sequences (star symbols) to
scribed into pre-CRISPR RNA (pre-crRNA) and subsequently initiate target recognition.

Phage

Cas1 Cas2

Class 1 New repeat New spacer


cas3 cas1 cas2
cascade

Class 2 L L
cas9 cas1 cas2 Acquisition
CRISPR array Expression
Genes encoding Cas proteins

pre-crRNA

Type I Type III Type IV


Class 1
Cas3 Cas10 Csf1
crRNAs

Type II Type V Type VI


Class 2 Interference
Cas9 Cas12 Cas13

Cas

Modern period. Following on from the proposal of CRISPR as an bacterium Streptococcus thermophilus. We discovered the exist-
immune system in prokaryotes came the first demonstration of its ence of CRISPR—still called SPIDR at that time—in this spe-
biological function. We had the privilege to be part of the team (at cies through a poster presenting the genome sequencing of strain
Danisco, a Danish food ingredient company now owned by DuPont) CNRZ1066 (ref. 17). Concurrently, we repeatedly came across
for which the various pieces required to solve the ‘CRISPR puzzle’ CRISPR loci while assembling the draft genomes of lactic acid bac-
were readily available: in silico datasets containing the genomes of teria commonly used in food manufacturing, including the pro-
bacteria encoding diverse CRISPR–Cas systems and the genomes of biotic Lactobacillus acidophilus 18. More interestingly, in late 2002,
bacteriophages able to infect them; the biological material, includ- we had access to the draft genome sequence for another strain of
ing bacterial isolates that carried active CRISPR–Cas systems and S.  thermophilus, LMD-9 (refs 17 and 19), an industrially relevant
lytic phages that could readily infect these hosts; and enough clues strain provided by Danisco, for which virulent phages were avail-
from the existing literature. able. We initially used CRISPR sequences to compare and contrast
This work started in the early 2000s during attempts to numerous S. thermophilus strains from the Danisco culture collec-
develop molecular methods for strain differentiation of the yogurt tion and observed the conservation of some spacers across groups

2 NATURE MICROBIOLOGY 2, 17092 (2017) | DOI: 10.1038/nmicrobiol.2017.92 | www.nature.com/naturemicrobiology


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
NATURE MICROBIOLOGY REVIEW ARTICLE
of strains. Remarkably, strains clustered in groups for which the
CRISPR genotypes correlated with phage resistance phenotypes.
Furthermore, we noticed that strains in our collection that had been Discovery of clustered repeats in E. coli5
isolated at various points in time (for instance, after phage chal-
lenges) did occasionally carry additional repeat–spacer sequences, Specific study of these clustered repeats in enterobacteria6

1990
suggesting phage-induced and time-dependent variability of
CRISPR sequences. Critically, we were able to sequence CRISPR loci
of industrial strains and their phage-resistant variants that had been
used and generated in the 1980s and 1990s, and then retroactively
shed light on the genomic basis for their phage resistance pheno-

Antiquity
type. By cross-referencing these data with phage genome sequences,
the link between the CRISPR genotype and the phage resistance Use of CRISPR for M. tuberculosis typing7
phenotype became noticeable, extending prior observations about

1995
an inverse correlation between the size of CRISPR loci and phage
Identification of CRISPRs in Haloferax9
propagation proportion in this species15. Next, we embarked on a
series of experiments using bacterial cultures and their lytic phages Description of CRISPRs in cyanobacteria10
to test the hypothesis that spacers get acquired during the genesis of
phage resistance. We were able to show that: (1) CRISPR loci have
the ability to uptake phage DNA as novel spacers during the adapta-
tion stage (Box 1); (2) altering the CRISPR spacer content is directly Use of CRISPR for S. pyogenes typing8
responsible for the gain, loss or exchange of phage resistance via

2000
addition, deletion or transplantation of CRISPR spacers, respec- Widespread occurrence of CRISPRs in bacteria and archaea11
tively; (3) cas genes are implicated in both the adaptation stage (csn2-
dependent integration of new spacers) and interference process DNA repair hypothesis for cas genes12
CRISPR name coined; description of cas genes2,3
Middle ages
(cas9-dependent resistance to lytic phages) (Box  1). These results
were published in the seminal 2007 study establishing CRISPR–Cas
as the adaptive immune system of bacteria20, essentially launching
the CRISPR field ‘renaissance’ (Fig. 1).
2005

While this may seem evident in hindsight, testing the hypoth- Spacers probably derive from foreign genetic elements13,14,15
esis that CRISPR–Cas systems are involved in phage resistance in
bacteria was not straightforward, given the need to carry out these Putative RNAi-based mechanism proposed16
experiments in a genetically challenging species, S. thermophilus, for CRISPR–Cas is an adaptive immune system against viruses
in S. thermophilus20
which molecular biology reagents and genome alteration tools were
Discovery of the proto-spacer adjacent motif in S. thermophilus31,32
somewhat limited at the time. Nevertheless, the S.  thermophilus
Modern period

Immunity is mediated by small CRISPR RNAs in E. coli28


model21 was surprisingly the first CRISPR–Cas system in which
Plasmid dsDNA is the target in S. epidermidis29
adaptation20, biogenesis of CRISPR RNAs (crRNAs)22,23 and inter- ssRNA is the target in P. furiosus38
2010

ference24 were characterized in native conditions; recently, these


Precise cleavage of target DNA by Cas9 in S. thermophilus24
steps were also characterized in Pseudomonas aeruginosa25–27. Discovery of the tracrRNA in S. pyogenes36
Within a year, two additional important discoveries were made: Heterologous transfer of a complete CRISPR–Cas system37
establishing that CRISPR arrays are transcribed and processed
Reprogrammable, Cas9-based DNA cleavage44,45
into crRNAs that guide Cas effector proteins, such as Cascade Cas9-based genome editing of human cells46,47 and bacteria48
(CRISPR-associated complex for antiviral defense)28; and that DNA
Contemporary period

is the target of CRISPR-encoded immunity, with this system being Cas9 structure109,111
Cas9-based eradication of latent HIV infection82
able to provide a barrier to plasmid transfer between bacteria29
2015

(Fig. 1; Box 1).
Cpf1 is a single RNA-guided nuclease134
Having established CRISPR as a DNA-encoded, RNA-mediated
and DNA-targeting immune system, a series of subsequent studies Discovery of anti-CRISPR proteins against Cas9 in N. meningitidis65
delved into the biochemical and genetic processes underpinning the
mechanism of action30 (Box 1). One critical example is the discovery
of the occurrence and conservation of a CRISPR motif flanking the
targeted phage sequences31–33, which was renamed the protospacer Figure 1 | CRISPR milestones and seminal discoveries. Select publications
adjacent motif (PAM)33. Furthermore, it was demonstrated that across the four listed eras are shown on the right. S. epidermidis,
Cas9 is an endonuclease that precisely generates blunt DNA cleavage Staphylococcus epidermidis; P. furiosus, Pyrococcus furiosus; N. meningitidis,
exactly 3 nucleotides from the 3ʹ edge of the protospacer sequence24. Neisseria meningitidis.
Cas9-mediated cleavage was also shown to be mediated by a seed
sequence, driving the specificity of targeting and intimacy of the
guide RNA:target DNA duplex 34,35. Subsequently in 2011, it was structures, as well as cleavage outcomes), especially with regard to
shown that in type II systems, an ancillary RNA, the trans-encoded the type of nucleic acid targeted as some CRISPR–Cas systems can
crRNA (tracrRNA), is necessary to create a dual RNA guide for Cas9 actually target RNA38,39. Currently, CRISPR–Cas systems are cat-
(ref. 36). This was also the year when it was first established that egorized into two main classes: class 1 has a multi-protein effector
CRISPR–Cas systems can be transferred to distant bacterial species, complex, whereas class 2 has a single effector protein. The classifi-
and reprogrammed to target plasmids and phages37 (Fig. 1). cation also includes 6 main types (I–VI) and over 19 subtypes40–43,
In light of these studies, it became increasingly clear that there based on the CRISPR and Cas machinery and the mode of action
are numerous and diverse CRISPR–Cas systems in nature that share involved (Box 2).
commonalities (DNA-encoded, RNA-mediated, nucleic acid target- The ‘modern period’ culminates with two milestone studies pub-
ing), but also carry out idiosyncrasies (various guide RNA types and lished in the summer of 2012 showing that Cas9 is an endonuclease

NATURE MICROBIOLOGY 2, 17092 (2017) | DOI: 10.1038/nmicrobiol.2017.92 | www.nature.com/naturemicrobiology 3


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEW ARTICLE NATURE MICROBIOLOGY

Box 2 | Origin and evolution of CRISPR–Cas systems.

Bacterial and archaeal genomes comprise an extraordinarily that employ the Cas1-like endonuclease as a recombinase, via an
diverse assortment of CRISPR–Cas systems that have been classi- integration/excision mechanism similar to that responsible for
fied into 2 main classes, 6 types and 19 subtypes42,142. In addition to the integration of new spacers into CRISPR loci143–145,147. Such a
the hyper-variable nature of CRISPR arrays, which is largely due to casposon, whose terminal inverted repeats might have evolved
the ability to capture novel spacers during the adaptation stage, the into CRISPR repeats, would have provided the adaptation mod-
CRISPR-associated cas gene pool also displays an extreme diver- ule of a CRISPR–Cas system ancestor in thermophilic Archaea
sification, probably resulting from numerous rearrangements and after fusion with a cas10-like nuclease gene originally involved in
horizontal transfer events over time40,42. Despite this high level of innate immunity, and other ancillary genes40,146. Further genetic
genetic complexity, most cas genes are organized into two main divergence and rearrangements probably gave rise to a diversity
modules that are respectively involved in adaptation and interfer- of class 1 systems, which then spread horizontally within prokary-
ence functions, and their detailed comparative analysis has led otic genomes. Subsequently, a distinct group of non-autonomous
to a simple and plausible scenario for the origin and evolution mobile elements named ISC (insertion sequences cas9-like, deriv-
of the Cas machinery, in which mobile elements played a pivotal ing from IS605 family transposons) potentially provided, by
role143–146. Indeed, cas1 homologues not associated with CRISPR– recombination happening at multiple and independent instances
Cas loci revealed a new prokaryotic superfamily of self-synthe- during evolution, the RuvC, HNH and/or HEPN endo(ribo)nucle-
sizing transposable elements (transposons) named casposons ase domains of the various class 2 CRISPR–Cas systems135,146,148.

that uses RuvC and HNH motifs to generate dual nicks in target resources. With a sense of déjà vu related to the discovery of restric-
DNA44, which can be reprogrammed using a single-guide RNA tion enzymes from restriction-modification systems69, bacterial
(sgRNA), mimicking the dual crRNA:tracrRNA complex 45 (Fig. 1). phage defence systems also ushered a new set of tools for molecular
These discoveries arguably set the stage for the repurposing of biology based on Cas nucleases, and a new array of CRISPR-based
CRISPR-derived molecular scalpels. applications for microbiologists. Indeed, the popularity of CRISPR
hinges on the portability of Cas nucleases, the programmability
Contemporary period. Within months, a number of laborato- of RNA guides, and the speed and efficiency with which CRISPR-
ries showed nearly concurrently that CRISPR–Cas9 molecular based machines can be manipulated.
machines can be repurposed to generate double-stranded breaks The many applications of CRISPR in eukaryotic organisms have
in DNA and lead to genome editing using DNA repair systems46–48 been extensively reviewed. For microbiologists, it is important to
(Fig. 1), giving rise to what has been termed the ‘CRISPR craze’49 note that both native and engineered systems can be repurposed or
and the democratization of genome editing 50,51. exploited broadly in bacteria and archaea.
Disruptive CRISPR-based technologies have since taken the
world by storm, based on a number of Cas-based molecular CRISPR-based genetic signatures. By essence, the CRISPR–Cas
machines (that is, Cas9, Cpf1, dCas9 (deactivated variants of Cas9)) adaptive immune system drives vaccination against invasive genetic
that enable genome editing, transcriptional control52–54 and epige- elements over time by capturing pieces of foreign DNA and insert-
netic alterations55. The many advantages and features of Cas pro- ing them as novel spacers into the CRISPR array, iteratively build-
teins (programmability, transferability, affordability, specificity and ing up a genetic record of immunization events20,31,70. These genetic
efficiency, amongst many) have empowered geneticists throughout ‘mug shots’ constitute a unique vaccination card that is hyper-vari-
the world and the phylogenetic spectrum to alter at will the genome able even amongst co-existing bacterial clones exposed to the same
and transcriptome of all genetically characterized species tested lytic virus71. Thus, they constitute individualized immunization
to date. records that serve as unique molecular signatures, and can be use-
Ironically, though CRISPR has been broadly adopted as a ful to genotype bacteria and track their presence and occurrence.
genome editing technology, there are no clustered regularly inter- Notably, due to the polarized mechanism of spacer acquisition at
spaced short palindromic repeats per  se in the tools used to alter the leader end of the CRISPR locus, leader-distal spacers provide
genetic sequences, but rather sgRNAs that direct Cas9 (or other) phylogenetic insights into ancestral origins of bacterial isolates,
nucleases for sequence-specific targeting of DNA. Notably, whereas while leader-proximal spacers provide insights into recent environ-
CRISPR is commonly perceived as a genome editor, Cas9 merely mental conditions of a bacterium (Fig. 2a). These loci thus consti-
cuts DNA (as initially intended when targeting phage DNA in bac- tute valuable genetic and epidemiological markers for the tracking
teria), while the actual editing is generated by the endogenous DNA of bacterial pathogens that are commonly problematic in the food
repair pathways. supply and responsible for infectious disease72,73, and analysis of
Current studies of CRISPR–Cas systems investigate the various these CRISPR-based molecular signatures could be explored by
aspects of the mechanism of action, with foci on novel immunity regulatory and monitoring agencies. Furthermore, unique com-
acquisition genetics and machinery 56–60, as well as elucidating the binations of spacers provide distinctive DNA fingerprints that
molecular and structural basis of interference by Cas nucleases61–63, enable natural genetic tagging and molecular tracking of bacterial
and its evolutionary implications, notably with regards to host– isolates, which have been explored in industry and basic research.
virus interactions25,64 and the impact of CRISPR immunity and For example, it was recently shown that adaptation can be repur-
targeting countermeasures27,65–68. posed to carry out molecular recording at a scalable level74 by prim-
ing CRISPR–Cas systems to acquire ‘donor’ DNA fragments as new
CRISPR applications in microbiology spacers into CRISPR arrays, yielding a unique molecular signature.
Traversing the valley of death in translational science equates to As our understanding and knowledge of microbiomes advance, it is
bridging the gap between seminal basic research discoveries and also possible to use hyper-variable CRISPR loci for high-resolution
their broad impactful applications and implementation. Often, genotyping of microbial populations64,75–77, with the ability to truly
success depends on the genesis of tools and technologies that ena- assess the genetic diversity of a given species78. Early efforts to use
ble users to perform valuable tasks with ease, speed and limited CRISPR to unravel genetic diversity in microbiomes have set the

4 NATURE MICROBIOLOGY 2, 17092 (2017) | DOI: 10.1038/nmicrobiol.2017.92 | www.nature.com/naturemicrobiology


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
NATURE MICROBIOLOGY REVIEW ARTICLE
a Sequence-based genotyping b Anti-infective systems
Unique to Shared across Ancestral Phage
one cluster two clusters spacer

L
L Cas9
Identical spacers Cluster 1
L Target
DNA
L crRNA
L
Iterative additions Cluster 2
L
L
cas9
L
Cluster 3 Genes encoding Cas proteins CRISPR array
L
Unique profiles Cluster 4
L
Cluster 5

c CRISPR-based antimicrobials d Genome editing

CRISPR CRISPR
Phage L Phage L
DNA DNA cas9 or dcas9

crRNA crRNA
Cas3 Cas3 Cas9

DNA repair
dCas9
cas3
Gene silencing
Genes encoding Cas proteins

Figure 2 | CRISPR-based applications. a, Exploitation of CRISPR sequences for sequence-based genotyping in bacteria. Unique combinations of spacer
sequences enable unravelling of phylogenetic relationships between strains, with conserved ancestral spacers at the leader-distal end (which can
be shared across genetic clusters), and polymorphic content at the leader-proximal end (which can distinguish strains even within a closely related
cluster). b, CRISPR as anti-infective systems. Native CRISPR–Cas systems on the bacterial chromosome enable DNA-encoded, RNA-mediated (crRNA),
DNA-targeting cleavage of invasive viral DNA by the endonuclease Cas9, providing adaptive immunity. c, CRISPR-based antimicrobials. Endogenous
CRISPR–Cas systems can be repurposed by delivering, via engineered phages, self-targeting spacers that direct the Cas nucleases for chromosome
targeting and degradation, leading to lethal DNA damage generated by the native Cas3 exonuclease. d, Microbial genome editing. Viral delivery of a
Cas9 nuclease and a guide RNA can generate a dsDNA break at the target chromosomal location, which can be edited using the endogenous DNA
repair pathway(s). Alternatively, a deactivated version of Cas9 (dCas9) that can still bind to target sequences but does not cleave DNA enables
sequence-specific binding of dCas9 to block transcription, repressing target gene expression.

stage and developed the necessary tools to dig into complex bacte- (dsDNA) stage in their life cycle, as well as viruses that may present
rial consortia and explore both bacterial composition and the inter- a health risk during xenotransplantations83.
play between hosts and viruses75,76,79,80. One originally unintended but practically desirable outcome of
effective phage sequence targeting by CRISPR is the selection of
A natural antiviral and anti-infective system. The basic biological escape phages31,84,85 through phage genome editing, essentially turn-
function of CRISPR–Cas immune systems can be used either natu- ing an escape mutation into an editing event by screening for muta-
rally or by engineering to build-up resistance against foreign mobile tions of the target sequence86,87. Practically, as phage genome editing
genetic elements (such as bacteriophages, plasmids or transposons), tools are relatively primitive, targeting a particular viral sequence
as well as to create a barrier against the uptake and dissemination using CRISPR can enable the user to screen and select for particu-
of undesirable genetic elements24 (such as antibiotic resistance lar mutations in the PAM or protospacer sequence that enable the
markers or toxin-encoding genes) (Fig. 2b). For example, iterative phage to escape CRISPR targeting, essentially equating to ‘natural’
vaccination events can be carried out to build-up broad-spectrum genome editing.
phage resistance in industrial bacterial strains, and such strategies
have been commercialized on a global scale for dairy cultures70,81. CRISPR-based antimicrobials. Whereas CRISPR–Cas systems
Additionally, there are also applications in generating beneficial originally evolved to protect the integrity of bacterial genomes,
bacteria that promote human, animal and plant health (see below). these systems can be repurposed for lethal self-targeting, culmi-
Moreover, these powerful antivirals have also been repurposed nating in bacterial elimination. Due to their lifestyle (single cells
to target viruses in eukaryotes, with noteworthy potential against with short lifespans and a high propensity and tolerance for genetic
HIV82 and possibly all viruses that include a double-stranded DNA fluidity) and limited access to complex and advanced DNA repair

NATURE MICROBIOLOGY 2, 17092 (2017) | DOI: 10.1038/nmicrobiol.2017.92 | www.nature.com/naturemicrobiology 5


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEW ARTICLE NATURE MICROBIOLOGY

pathways (compared to their eukaryotic counterparts), bacteria are these reports highlight the interplay of homology and convergence
particularly susceptible to genomic DNA damage. Consequently, in the evolution of these effector nucleases, and illustrate how
the primary outcome of CRISPR-encoded self-targeting in bacteria nucleic acid targeting domains and catalytic sites have evolved and
is actually death88,89, rather than genome editing. This can be readily specialized to effectively target various types of invasive molecules
harnessed to drive programmable and sequence-specific bacterial in a nimble and diversified manner, perhaps reflecting the diversity
cell death in microbial populations, especially when using the potent of invaders and predators. Importantly, the genesis of orthogonal
and damaging type I CRISPR–Cas systems90 (Fig. 2c). Recently, sev- systems will enable the concurrent use and multiplexing of various
eral studies have documented the selective antimicrobial potential applications, such as metabolic rerouting by simultaneous upregula-
of CRISPR-based antimicrobials91,92, and opened new avenues for tion of specific pathways and downregulation of others124–126.
the use of CRISPR-mediated self-targeting for the selective removal The ability to repurpose native systems in bacteria also enables
of bacterial genotypes of interest (and the modulation of the com- the screening of some rare natural variations, such as the excision
position of microbiomes) in preventing infectious disease, and also of expandable genomic islands127. Furthermore, CRISPR can be
reformulating bacterial consortia to promote the health of animals readily used in combination with single-strand DNA recombineer-
and plants91. Strategically selecting the most damaging CRISPR–Cas ing (recombination-mediated genetic engineering) technologies to
systems will promote lethality, and the naturally dominant type I quickly and efficiently modify the genomic content of industrial
system, with the signature exonuclease Cas3, holds much potential workhorses and model microorganisms, such as lactobacilli, that
in the genesis of extensive DNA damage in bacteria93–98. Indeed, are widely used as commercial probiotics128. These technologies
generating more extensive DNA damage would be more challenging have the potential to be implemented on a broad scale in all bacte-
to repair, or costlier to forego, consequently enabling more potent ria of industrial interest, encompassing biomanufacturing for food,
eradication of a bacterial population. Furthermore, the widespread agricultural and biotechnological applications129,130.
nature of type I CRISPR–Cas systems99, notably in pathogenic bac-
teria such as Salmonella, E.  coli or Clostridium difficile, opens the The CRISPR future is now
option to repurpose endogenous CRISPR effector nucleases to The current pace and scale at which CRISPR-based technologies are
selectively eliminate these pathogens. being used is undeniably remarkable and is broadly perceived as the
tip of the iceberg, with unabating momentum in terms of science
Next-generation genome editing. As a nod to the repurposing of and frenetic business interest. Indeed, from a scientific standpoint, it
microbial systems for genome editing in eukaryotes, microbiologists appears our imagination is the sole limiting factor for what CRISPR-
have recently repurposed the sgRNA:Cas9 technology for genome based technologies can do, in terms of both applications and imple-
editing of bacteria (Fig.  2d). Indeed, the various genome editing mentable species. The literature and citation quantitative patterns
applications pioneered in human and animal cells have recently are compelling and reflect rapid and nearly ubiquitous adoption
been transferred (back) to bacteria to carry out genome editing of this technology by most investigators studying genetics4,49.  The
and transcriptional control, as well as genome-wide screens100. accolades for various CRISPR pioneers have been humbling, though
Self-targeting can be used to alter the genetic content of bacte- arguably biased, and not always representative of the humble and
ria quickly and scalably, with exquisite specificity (since bacterial diverse bacterial roots of these fantastic molecular machines131,132.
genomes are so much smaller than their eukaryotic counterparts, From a business standpoint, the translational potential of this tech-
with much less repeated DNA) and efficiency (since prokaryotic nology for both gene therapies and agricultural breeding have gen-
molecular machines work so well in bacteria). Furthermore, dCas9, erated sizeable investments and impressive strategic partnerships
in which the RuvC and HNH nickase domains have been mutated, in fields that span the pharmaceutical, agricultural, food and bio-
have the ability to alter transcription patterns by either blocking technology industries. Of course, this enthusiasm is accompanied
RNA polymerase (CRISPR interference, CRISPRi) or by triggering by commensurately high expectations that may prove difficult to
gene expression (CRISPR activation, CRISPRa), enabling rerout- live up to. Unfortunately, this has occasionally come at the cost of
ing of transcriptional pathways52–54,101. Again, microbiologists may controversial coverage, such as journalistic sensationalism and an
either exploit engineered, or repurpose endogenous, CRISPR–Cas epic intellectual property battle, but this reflects the appetite for,
systems to drive transcriptional control102,103. While the popularity and potential of, this technology. Furthermore, this also creates a
and success of the CRISPR-fueled genome-editing revolution hinge tangible need to safely and responsibly manage the development
on the appetite built by RNAi and first-generation (ZFN-, TALEN- and implementation of CRISPR-based technologies with immedi-
and meganuclease-based) technologies104, the microbial genome ate implications for regulatory processes, ethical considerations,
editing movement will rely on the adaptation of eukaryote-focused and public relations133, which is sensitive in light of the stakes, per-
tools and the continued progress of microbiome-focused research, ception and previous shortcomings of GMO (genetically modified
and identification of both the detrimental pathogens that need to organism)-associated technologies.
be eradicated, and the beneficial microorganisms that need to be In many ways, the CRISPR story illustrates how science often
harnessed and improved. works, with a fitting reminder that the basic microbiological pro-
Technological enhancements driven by an expanded exploita- cesses driving the survival of bacteria often give rise to valuable
tion of diverse CRISPR–Cas systems105–107, as well as the develop- molecular tools, enzymes and technologies, together with the his-
ment of engineered variants with improved functionalities based on torical involvement of a plethora of unsung heroes whom selflessly
structural insights108–113, set the stage for ever-increasing expansion advance science for the benefits of humankind. There is also a sense
of the CRISPR–Cas toolbox. For example, rational engineering has of déjà vu with the serendipitous timing and assembly of dispersed
been implemented to alter PAM recognition and optimize efficiency clues via a community-based scientific process with sometimes
and specificity of Cas9 nucleases114–117. More recently, several studies parallel efforts and convergent foci. The collaborative, diverse and
have focused on determining the structures of other class 2 effector multidisciplinary nature of the early CRISPR community built—
nucleases, including C2c1 and C2c2 (Cas13), unravelling the tar- a decade ago—the strong momentum that enabled the progress
get recognition and cleavage mechanisms of action for both RNA- observed ever since.
guided DNA and RNA targeting, respectively 118–123. These studies The recent past illustrates the pace at which the field is advanc-
may open new avenues to use Cas-based nucleases for RNA target- ing and the CRISPR literature exploding. Several recent studies
ing and detection, and enable a deeper understanding of the origin have reminded us that the bacterial dark matter still holds intrigu-
and diversification of CRISPR–Cas systems (Box  2). Remarkably, ing novel systems that can be repurposed for various applications,

6 NATURE MICROBIOLOGY 2, 17092 (2017) | DOI: 10.1038/nmicrobiol.2017.92 | www.nature.com/naturemicrobiology


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
NATURE MICROBIOLOGY REVIEW ARTICLE
such as a Cpf1 (Cas12) (ref. 134) and C2c1-2 (Cas13) (ref. 135), as 16. Makarova, K. S., Grishin, N. V., Shabalina, S. A., Wolf, Y. I. & Koonin, E. V. A
well as recently unearthed distant homologues of Cas9 and Cas12, putative RNA-interference-based immune system in prokaryotes: computational
analysis of the predicted enzymatic machinery, functional analogies with
namely CasX and CasY (ref. 136). These molecular machines con- eukaryotic RNAi, and hypothetical mechanisms of action. Biol. Direct 1, 7 (2006).
tinue to diversify the target nucleic acid and cleavage types, as well 17. Klaenhammer, T. et al. Discovering lactic acid bacteria by genomics. Antonie
as the targetable landscape of sequences using PAM diversity 137. Van Leeuwenhoek 82, 29–58 (2002).
Additionally, novel effectors, such as the anti-CRISPR proteins, can 18. Altermann, E. et al. Complete genome sequence of the probiotic lactic
be harnessed to control the activity of Cas enzymes either in nature acid bacterium Lactobacillus acidophilus NCFM. Proc. Natl Acad. Sci. USA
to evade CRISPR immunity 27, or in vitro to control the activity of 102, 3906–3912 (2005).
19. Makarova, K. et al. Comparative genomics of the lactic acid bacteria. Proc.
CRISPR-based technologies65. As always, many outstanding ques- Natl Acad. Sci. USA 103, 15611–15616 (2006).
tions still remain, notably the presence of uncharacterized CRISPR– 20. Barrangou, R. et al. CRISPR provides acquired resistance against viruses in
Cas immune systems in nature and the rationale for their absence in prokaryotes. Science 315, 1709–1712 (2007).
more than half of known bacterial species, as well as their biological 21. Horvath, P. & Barrangou, R. CRISPR/Cas, the immune system of bacteria and
roles beyond adaptive immunity. Actually, in addition to providing archaea. Science 327, 167–170 (2010).
adaptive immunity, CRISPR systems have also been implicated in 22. Carte, J. et al. The three major types of CRISPR-Cas systems function
independently in CRISPR RNA biogenesis in Streptococcus thermophilus. Mol.
biofilm formation and transcriptional control in bacteria, and there Microbiol. 93, 98–112 (2014).
may be additional functional roles138–141. 23. Karvelis, T. et al. crRNA and tracrRNA guide Cas9-mediated DNA
Certainly, this past decade has been filled with milestone interference in Streptococcus thermophilus. RNA Biol. 10, 841–851 (2013).
discoveries unravelling the genetic and biochemical basis of 24. Garneau, J. E. et al. The CRISPR/Cas bacterial immune system cleaves
CRISPR-based immunity in bacteria, as well as the repurposing bacteriophage and plasmid DNA. Nature 468, 67–71 (2010).
of native and engineered systems for a plethora of applications, 25. van Houte, S. et al. The diversity-generating benefits of a prokaryotic adaptive
immune system. Nature 532, 385–388 (2016).
and we already look forward to a promising upcoming decade of 26. Haurwitz, R. E., Sternberg, S. H. & Doudna, J. A. Csy4 relies on an
additional advances. unusual catalytic dyad to position and cleave CRISPR RNA. EMBO J.
31, 2824–2832 (2012).
Received 23 December 2016; accepted 5 May 2017; 27. Bondy-Denomy, J., Pawluk, A., Maxwell, K. L. & Davidson, A. R.
published 5 June 2017 Bacteriophage genes that inactivate the CRISPR/Cas bacterial immune system.
Nature 493, 429–432 (2013).
References 28. Brouns, S. J. et al. Small CRISPR RNAs guide antiviral defense in prokaryotes.
1. Labrie, S. J., Samson, J. E. & Moineau, S. Bacteriophage resistance Science 321, 960–964 (2008).
mechanisms. Nat. Rev. Microbiol. 8, 317–327 (2010). 29. Marraffini, L. A. & Sontheimer, E. J. CRISPR interference limits
2. Jansen, R., van Embden, J. D., Gaastra, W. & Schouls, L. M. Identification horizontal gene transfer in staphylococci by targeting DNA. Science
of a novel family of sequence repeats among prokaryotes. OMICS 322, 1843–1845 (2008).
6, 23–33 (2002). 30. Marraffini, L. A. CRISPR-Cas immunity in prokaryotes. Nature
3. Jansen, R., Embden, J. D., Gaastra, W. & Schouls, L. M. Identification of 526, 55–61 (2015).
genes that are associated with DNA repeats in prokaryotes. Mol. Microbiol. 31. Deveau, H. et al. Phage response to CRISPR-encoded resistance in
43, 1565–1575 (2002). Streptococcus thermophilus. J. Bacteriol. 190, 1390–1400 (2008).
4. Barrangou, R. & Doudna, J. A. Applications of CRISPR technologies in 32. Horvath, P. et al. Diversity, activity, and evolution of CRISPR loci in
research and beyond. Nat. Biotechnol. 34, 933–941 (2016). Streptococcus thermophilus. J. Bacteriol. 190, 1401–1412 (2008).
5. Ishino, Y., Shinagawa, H., Makino, K., Amemura, M. & Nakata, A. Nucleotide 33. Mojica, F. J., Diez-Villasenor, C., Garcia-Martinez, J. & Almendros, C. Short
sequence of the iap gene, responsible for alkaline phosphatase isozyme motif sequences determine the targets of the prokaryotic CRISPR defence
conversion in Escherichia coli, and identification of the gene product. system. Microbiology 155, 733–740 (2009).
J. Bacteriol. 169, 5429–5433 (1987). 34. Semenova, E. et al. Interference by clustered regularly interspaced short
6. Nakata, A., Amemura, M. & Makino, K. Unusual nucleotide arrangement palindromic repeat (CRISPR) RNA is governed by a seed sequence. Proc. Natl
with repeated sequences in the Escherichia coli K-12 chromosome. J. Bacteriol. Acad. Sci USA 108, 10098–10103 (2011).
171, 3553–3556 (1989). 35. Wiedenheft, B. et al. RNA-guided complex from a bacterial immune system
7. Groenen, P. M., Bunschoten, A. E., van Soolingen, D. & van Embden, J. D. enhances target recognition through seed sequence interactions. Proc. Natl
Nature of DNA polymorphism in the direct repeat cluster of Mycobacterium Acad. Sci. USA 108, 10092–10097 (2011).
tuberculosis; application for strain differentiation by a novel typing method. 36. Deltcheva, E. et al. CRISPR RNA maturation by trans-encoded small RNA and
Mol. Microbiol. 10, 1057–1065 (1993). host factor RNase III. Nature 471, 602–607 (2011).
8. Hoe, N. et al. Rapid molecular genetic subtyping of serotype M1 group A 37. Sapranauskas, R. et al. The Streptococcus thermophilus CRISPR/Cas system
Streptococcus strains. Emerg. Infect. Dis. 5, 254–263 (1999). provides immunity in Escherichia coli. Nucleic Acids Res 39, 9275–9282 (2011).
9. Mojica, F. J., Ferrer, C., Juez, G. & Rodriguez-Valera, F. Long stretches of short 38. Hale, C. R. et al. RNA-guided RNA cleavage by a CRISPR RNA-Cas protein
tandem repeats are present in the largest replicons of the Archaea Haloferax complex. Cell 139, 945–956 (2009).
mediterranei and Haloferax volcanii and could be involved in replicon 39. Jiang, W., Samai, P. & Marraffini, L. A. Degradation of phage transcripts by
partitioning. Mol. Microbiol. 17, 85–93 (1995). CRISPR-associated RNases enables type III CRISPR-Cas immunity. Cell
10. Masepohl, B., Gorlitz, K. & Bohme, H. Long tandemly repeated repetitive 164, 710–721 (2016).
(LTRR) sequences in the filamentous cyanobacterium Anabaena sp. PCC 40. Makarova, K. S., Aravind, L., Wolf, Y. I. & Koonin, E. V. Unification of
7120. Biochim. Biophys. Acta 1307, 26–30 (1996). Cas protein families and a simple scenario for the origin and evolution of
11. Mojica, F. J., Diez-Villasenor, C., Soria, E. & Juez, G. Biological significance of CRISPR-Cas systems. Biol. Direct 6, 38 (2011).
a family of regularly spaced repeats in the genomes of Archaea, Bacteria and 41. Makarova, K. S. et al. Evolution and classification of the CRISPR-Cas systems.
mitochondria. Mol. Microbiol. 36, 244–246 (2000). Nat. Rev. Microbiol. 9, 467–477 (2011).
12. Makarova, K. S., Aravind, L., Grishin, N. V., Rogozin, I. B. & Koonin, E. V. A 42. Makarova, K. S. et al. An updated evolutionary classification of CRISPR-Cas
DNA repair system specific for thermophilic Archaea and bacteria predicted systems. Nat. Rev. Microbiol. 13, 722–736 (2015).
by genomic context analysis. Nucleic Acids Res. 30, 482–496 (2002). 43. Makarova, K. S. & Koonin, E. V. Annotation and classification of CRISPR-Cas
13. Mojica, F. J., Diez-Villasenor, C., Garcia-Martinez, J. & Soria, E. Intervening systems. Methods Mol. Biol. 1311, 47–75 (2015).
sequences of regularly spaced prokaryotic repeats derive from foreign genetic 44. Gasiunas, G., Barrangou, R., Horvath, P. & Siksnys, V. Cas9–crRNA
elements. J. Mol. Evol. 60, 174–182 (2005). ribonucleoprotein complex mediates specific DNA cleavage for adaptive
14. Pourcel, C., Salvignol, G. & Vergnaud, G. CRISPR elements in Yersinia immunity in bacteria. Proc. Natl Acad. Sci. USA 109, E2579–E2586 (2012).
pestis acquire new repeats by preferential uptake of bacteriophage DNA, 45. Jinek, M. et al. A programmable dual-RNA-guided DNA endonuclease in
and provide additional tools for evolutionary studies. Microbiology adaptive bacterial immunity. Science 337, 816–821 (2012).
151, 653–663 (2005). 46. Cong, L. et al. Multiplex genome engineering using CRISPR/Cas systems.
15. Bolotin, A., Quinquis, B., Sorokin, A. & Ehrlich, S. D. Clustered regularly Science 339, 819–823 (2013).
interspaced short palindrome repeats (CRISPRs) have spacers of 47. Mali, P. et al. RNA-guided human genome engineering via Cas9. Science
extrachromosomal origin. Microbiology 151, 2551–2561 (2005). 339, 823–826 (2013).

NATURE MICROBIOLOGY 2, 17092 (2017) | DOI: 10.1038/nmicrobiol.2017.92 | www.nature.com/naturemicrobiology 7


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEW ARTICLE NATURE MICROBIOLOGY
48. Jiang, W., Bikard, D., Cox, D., Zhang, F. & Marraffini, L. A. RNA-guided 80. Andersson, A. F. & Banfield, J. F. Virus population dynamics and acquired virus
editing of bacterial genomes using CRISPR-Cas systems. Nat. Biotechnol. resistance in natural microbial communities. Science 320, 1047–1050 (2008).
31, 233–239 (2013). 81. Barrangou, R. & Horvath, P. CRISPR: new horizons in phage resistance and
49. Pennisi, E. The CRISPR craze. Science 341, 833–836 (2013). strain identification. Annu. Rev. Food Sci. Technol. 3, 143–162 (2012).
50. Barrangou, R. & May, A. P. Unraveling the potential of CRISPR-Cas9 for gene 82. Hu, W. et al. RNA-directed gene editing specifically eradicates
therapy. Expert Opin. Biol. Ther. 15, 311–314 (2015). latent and prevents new HIV-1 infection. Proc. Natl Acad. Sci. USA
51. Ledford, H. CRISPR, the disruptor. Nature 522, 20–24 (2015). 111, 11461–11466 (2014).
52. Gilbert, L. A. et al. Genome-Scale CRISPR-mediated control of gene 83. Yang, L. et al. Genome-wide inactivation of porcine endogenous retroviruses
repression and activation. Cell 159, 647–661 (2014). (PERVs). Science 350, 1101–1104 (2015).
53. Gilbert, L. A. et al. CRISPR-mediated modular RNA-guided regulation of 84. Sun, C. L. et al. Phage mutations in response to CRISPR diversification in a
transcription in eukaryotes. Cell 154, 442–451 (2013). bacterial population. Environ. Microbiol. 15, 463–470 (2013).
54. Qi, L. S. et al. Repurposing CRISPR as an RNA-guided platform for sequence- 85. Paez-Espino, D. et al. CRISPR immunity drives rapid phage genome evolution
specific control of gene expression. Cell 152, 1173–1183 (2013). in Streptococcus thermophilus. mBio 6, e00262-15 (2015).
55. Hilton, I. B. et al. Epigenome editing by a CRISPR-Cas9-based 86. Box, A. M., McGuffie, M. J., O’Hara, B. J. & Seed, K. D. Functional analysis
acetyltransferase activates genes from promoters and enhancers. of bacteriophage immunity through a type I-E CRISPR-Cas System in Vibrio
Nat. Biotechnol. 33, 510–517 (2015). cholerae and its application in bacteriophage genome engineering. J. Bacteriol.
56. Heler, R., Marraffini, L. A. & Bikard, D. Adapting to new threats: the 198, 578–590 (2015).
generation of memory by CRISPR-Cas immune systems. Mol. Microbiol. 87. Martel, B. & Moineau, S. CRISPR-Cas: an efficient tool for
93, 1–9 (2014). genome engineering of virulent bacteriophages. Nucleic Acids Res.
57. Heler, R. et al. Cas9 specifies functional viral targets during CRISPR-Cas 42, 9504–9513 (2014).
adaptation. Nature 519, 199–202 (2015). 88. Vercoe, R. B. et al. Cytotoxic chromosomal targeting by CRISPR/Cas systems
58. Levy, A. et al. CRISPR adaptation biases explain preference for acquisition of can reshape bacterial genomes and expel or remodel pathogenicity islands.
foreign DNA. Nature 520, 505–510 (2015). PLoS Genet. 9, e1003454 (2013).
59. Nunez, J. K., Harrington, L. B., Kranzusch, P. J., Engelman, A. N. & Doudna, 89. Edgar, R. & Qimron, U. The Escherichia coli CRISPR system protects
J. A. Foreign DNA capture during CRISPR-Cas adaptive immunity. Nature from λ lysogenization, lysogens, and prophage induction. J. Bacteriol.
527, 535–538 (2015). 192, 6291–6294 (2010).
60. Nunez, J. K., Lee, A. S., Engelman, A. & Doudna, J. A. Integrase-mediated 90. Gomaa, A. A. et al. Programmable removal of bacterial strains by use of
spacer acquisition during CRISPR-Cas adaptive immunity. Nature 519, genome-targeting CRISPR-Cas systems. mBio 5, e00928-13 (2014).
193–198 (2015). 91. Bikard, D. et al. Exploiting CRISPR-Cas nucleases to produce sequence-
61. Sternberg, S. H., LaFrance, B., Kaplan, M. & Doudna, J. A. Conformational specific antimicrobials. Nat. Biotechnol. 32, 1146–1150 (2014).
control of DNA target cleavage by CRISPR-Cas9. Nature 527, 110–113 (2015). 92. Citorik, R. J., Mimee, M. & Lu, T. K. Sequence-specific
62. Sternberg, S. H., Redding, S., Jinek, M., Greene, E. C. & Doudna, J. A. DNA antimicrobials using efficiently delivered RNA-guided nucleases. Nat.
interrogation by the CRISPR RNA-guided endonuclease Cas9. Nature Biotechnol. 32, 1141–1145 (2014).
507, 62–67 (2014). 93. Gong, B. et al. Molecular insights into DNA interference by CRISPR-
63. Szczelkun, M. D. et al. Direct observation of R-loop formation by single associated nuclease-helicase Cas3. Proc. Natl Acad. Sci. USA
RNA-guided Cas9 and Cascade effector complexes. Proc. Natl Acad. Sci. USA 111, 16359–16364 (2014).
111, 9798–9803 (2014). 94. Hochstrasser, M. L. et al. CasA mediates Cas3-catalyzed target degradation
64. Sun, C. L., Thomas, B. C., Barrangou, R. & Banfield, J. F. Metagenomic during CRISPR RNA-guided interference. Proc. Natl Acad. Sci. USA
reconstructions of bacterial CRISPR loci constrain population histories. 111, 6618–6623 (2014).
ISME J. 10, 858–870 (2016). 95. Huo, Y. et al. Structures of CRISPR Cas3 offer mechanistic insights into
65. Pawluk, A. et al. Naturally occurring off-switches for CRISPR-Cas9. Cell Cascade-activated DNA unwinding and degradation. Nat. Struct. Mol. Biol.
167, 1829–1838 (2016). 21, 771–777 (2014).
66. Pawluk, A. et al. Inactivation of CRISPR-Cas systems by anti-CRISPR proteins 96. Sinkunas, T. et al. Cas3 is a single-stranded DNA nuclease and ATP-
in diverse bacterial species. Nat. Microbiol. 1, 16085 (2016). dependent helicase in the CRISPR/Cas immune system. EMBO J.
67. Seed, K. D., Lazinski, D. W., Calderwood, S. B. & Camilli, A. A bacteriophage 30, 1335–1342 (2011).
encodes its own CRISPR/Cas adaptive response to evade host innate 97. Sinkunas, T. et al. In vitro reconstitution of Cascade-mediated CRISPR
immunity. Nature 494, 489–491 (2013). immunity in Streptococcus thermophilus. EMBO J. 32, 385–394 (2013).
68. Bondy-Denomy, J. et al. Multiple mechanisms for CRISPR-Cas inhibition by 98. Westra, E. R. et al. CRISPR immunity relies on the consecutive binding and
anti-CRISPR proteins. Nature 526, 136–139 (2015). degradation of negatively supercoiled invader DNA by Cascade and Cas3.
69. Loenen, W. A., Dryden, D. T., Raleigh, E. A., Wilson, G. G. & Murray, N. E. Mol. Cell 46, 595–605 (2012).
Highlights of the DNA cutters: a short history of the restriction enzymes. 99. Grissa, I., Vergnaud, G. & Pourcel, C. The CRISPRdb database and tools to
Nucleic Acids Res. 42, 3–19 (2014). display CRISPRs and to generate dictionaries of spacers and repeats. BMC
70. Barrangou, R. et al. Genomic impact of CRISPR immunization against Bioinformatics 8, 172 (2007).
bacteriophages. Biochem. Soc. Trans. 41, 1383–1391 (2013). 100. Peters, J. M. et al. A comprehensive, CRISPR-based functional analysis of
71. Paez-Espino, D. et al. Strong bias in the bacterial CRISPR elements that confer essential genes in bacteria. Cell 165, 1493–1506 (2016).
immunity to phage. Nat. Commun. 4, 1430 (2013). 101. Larson, M. H. et al. CRISPR interference (CRISPRi) for sequence-specific
72. Shariat, N. et al. Subtyping of Salmonella enterica Serovar Newport outbreak control of gene expression. Nat. Protoc. 8, 2180–2196 (2013).
isolates by CRISPR-MVLST and determination of the relationship between 102. Bikard, D. et al. Programmable repression and activation of bacterial gene
CRISPR-MVLST and PFGE results. J. Clin. Microbiol. 51, 2328–2336 (2013). expression using an engineered CRISPR-Cas system. Nucleic Acids Res.
73. Barrangou, R. & Dudley, E. G. CRISPR-based typing and next-generation 41, 7429–7437 (2013).
tracking technologies. Annu. Rev. Food Sci. Technol. 7, 395–411 (2016). 103. Luo, M. L., Mullis, A. S., Leenay, R. T. & Beisel, C. L. Repurposing endogenous
74. Shipman, S. L., Nivala, J., Macklis, J. D. & Church, G. M. Molecular recordings type I CRISPR-Cas systems for programmable gene repression. Nucleic Acids
by directed CRISPR spacer acquisition. Science http://dx.doi.org/10.1126/ Res. 43, 674–681 (2015).
science.aaf1175 (2016). 104. Barrangou, R. et al. Advances in CRISPR-Cas9 genome engineering: lessons
75. Tyson, G. W. & Banfield, J. F. Rapidly evolving CRISPRs implicated in learned from RNA interference. Nucleic Acids Res. 43, 3407–3419 (2015).
acquired resistance of microorganisms to viruses. Environ. Microbiol. 105. Makarova, K. S., Zhang, F. & Koonin, E. V. Snapshot: class 1 CRISPR-Cas
10, 200–207 (2008). systems. Cell 168, 946–946.e1 (2017).
76. Pride, D. T. et al. Analysis of streptococcal CRISPRs from human saliva reveals 106. Makarova, K. S., Zhang, F. & Koonin, E. V. Snapshot: class 2 CRISPR-Cas
substantial sequence diversity within and between subjects over time. Genome systems. Cell 168, 328–328.e1 (2017).
Res. 21, 126–136 (2011). 107. Barrangou, R. Diversity of CRISPR-Cas immune systems and molecular
77. Weinberger, A. D. et al. Persisting viral sequences shape microbial CRISPR- machines. Genome Biol. 16, 247 (2015).
based immunity. PLoS Comput. Biol. 8, e1002475 (2012). 108. Nishimasu, H. et al. Crystal structure of Staphylococcus aureus Cas9. Cell
78. Briner, A. E. & Barrangou, R. Deciphering and shaping bacterial diversity 162, 1113–1126 (2015).
through CRISPR. Curr. Opin. Microbiol. 31, 101–108 (2016). 109. Nishimasu, H. et al. Crystal structure of Cas9 in complex with guide RNA and
79. Heidelberg, J. F., Nelson, W. C., Schoenfeld, T. & Bhaya, D. Germ warfare in a target DNA. Cell 156, 935–949 (2014).
microbial mat community: CRISPRs provide insights into the co-evolution of 110. Yamano, T. et al. Crystal Structure of Cpf1 in complex with guide RNA and
host and viral genomes. PLoS ONE 4, e4169 (2009). target DNA. Cell 165, 949–962 (2016).

8 NATURE MICROBIOLOGY 2, 17092 (2017) | DOI: 10.1038/nmicrobiol.2017.92 | www.nature.com/naturemicrobiology


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
NATURE MICROBIOLOGY REVIEW ARTICLE
111. Anders, C., Niewoehner, O., Duerst, A. & Jinek, M. Structural basis of 137. Leenay, R. T. et al. Identifying and visualizing functional PAM diversity across
PAM-dependent target DNA recognition by the Cas9 endonuclease. Nature CRISPR-Cas systems. Mol. Cell 62, 137–147 (2016).
513, 569–573 (2014). 138. Sampson, T. R. et al. A CRISPR-Cas system enhances envelope integrity
112. Jinek, M. et al. Structures of Cas9 endonucleases reveal RNA-mediated mediating antibiotic resistance and inflammasome evasion. Proc. Natl Acad.
conformational activation. Science 343, 1247997 (2014). Sci. USA 111, 11163–11168 (2014).
113. Anders, C., Bargsten, K. & Jinek, M. Structural plasticity of PAM recognition 139. Sampson, T. R., Saroj, S. D., Llewellyn, A. C., Tzeng, Y. L. & Weiss, D. S. A
by engineered variants of the RNA-guided endonuclease Cas9. Mol. Cell CRISPR/Cas system mediates bacterial innate immune evasion and virulence.
61, 895–902 (2016). Nature 497, 254–257 (2013).
114. Slaymaker, I. M. et al. Rationally engineered Cas9 nucleases with improved 140. Louwen, R. et al. A novel link between Campylobacter jejuni bacteriophage
specificity. Science 351, 84–88 (2016). defence, virulence and Guillain-Barre syndrome. Eur. J. Clin. Microbiol. Infect.
115. Kleinstiver, B. P. et al. High-fidelity CRISPR-Cas9 nucleases with no detectable Dis. 32, 207–226 (2013).
genome-wide off-target effects. Nature 529, 490–495 (2016). 141. Louwen, R., Staals, R. H., Endtz, H. P., van Baarlen, P. & van der Oost, J. The
116. Kleinstiver, B. P. et al. Broadening the targeting range of Staphylococcus role of CRISPR-Cas systems in virulence of pathogenic bacteria. Microbiol.
aureus CRISPR-Cas9 by modifying PAM recognition. Nat. Biotechnol. Mol. Biol. Rev. 78, 74–88 (2014).
33, 1293–1298 (2015). 142. Shmakov, S. et al. Diversity and evolution of class 2 CRISPR-Cas systems. Nat.
117. Kleinstiver, B. P. et al. Engineered CRISPR-Cas9 nucleases with altered PAM Rev. Microbiol. 15, 169–182 (2017).
specificities. Nature 523, 481–485 (2015). 143. Koonin, E. V. & Krupovic, M. Evolution of adaptive immunity from
118. Yang, H., Gao, P., Rajashankar, K. R. & Patel, D. J. PAM-dependent target transposable elements combined with innate immune systems. Nat. Rev.
DNA recognition and cleavage by C2c1 CRISPR-Cas endonuclease. Cell Genet. 16, 184–192 (2015).
167, 1814–1828 (2016). 144. Krupovic, M., Makarova, K. S., Forterre, P., Prangishvili, D. & Koonin, E. V.
119. Liu, L. et al. C2c1-sgRNA complex structure reveals RNA-guided DNA Casposons: a new superfamily of self-synthesizing DNA transposons at the
cleavage mechanism. Mol. Cell 65, 310–322 (2017). origin of prokaryotic CRISPR-Cas immunity. BMC Biol. 12, 36 (2014).
120. Liu, L. et al. Two distant catalytic sites are responsible for C2c2 RNase 145. Krupovic, M., Shmakov, S., Makarova, K. S., Forterre, P. & Koonin, E. V.
activities. Cell 168, 121–134 (2017). Recent mobility of casposons, self-synthesizing transposons at the origin of
121. Lewis, K. M. & Ke, A. Building the class 2 CRISPR-Cas arsenal. Mol. Cell the CRISPR-Cas immunity. Genome Biol. Evol. 8, 375–386 (2016).
65, 377–379 (2017). 146. Mohanraju, P. et al. Diverse evolutionary roots and mechanistic variations of
122. Abudayyeh, O. O. et al. C2c2 is a single-component programmable RNA- the CRISPR-Cas systems. Science 353, aad5147 (2016).
guided RNA-targeting CRISPR effector. Science 353, aaf5573 (2016). 147. Makarova, K. S., Wolf, Y. I. & Koonin, E. V. The basic building
123. East-Seletsky, A. et al. Two distinct RNase activities of CRISPR-C2c2 enable blocks and evolution of CRISPR-CAS systems. Biochem. Soc. Trans.
guide-RNA processing and RNA detection. Nature 538, 270–273 (2016). 41, 1392–1400 (2013).
124. Briner, A. E. et al. Guide RNA functional modules direct Cas9 activity and 148. Kapitonov, V. V., Makarova, K. S. & Koonin, E. V. ISC, a novel group of
orthogonality. Mol. Cell 56, 333–339 (2014). bacterial and archaeal DNA transposons that encode Cas9 homologs.
125. Esvelt, K. M. et al. Orthogonal Cas9 proteins for RNA-guided gene regulation J. Bacteriol. 198, 797–807 (2015).
and editing. Nat. Methods 10, 1116–1121 (2013).
126. Fonfara, I. et al. Phylogeny of Cas9 determines functional exchangeability of Acknowledgements
dual-RNA and Cas9 among orthologous type II CRISPR-Cas systems. Nucleic R.B. is supported by funds from North Carolina State University and the North Carolina
Acids Res. 42, 2577–2590 (2014). Ag. Foundation. The authors would like to thank their colleagues and collaborators for
127. Selle, K., Klaenhammer, T. R. & Barrangou, R. CRISPR-based screening their contributions and insights, and for having the privilege to share the CRISPR journey.
of genomic island excision events in bacteria. Proc. Natl Acad. Sci. USA
112, 8076–8081 (2015).
128. Oh, J. H. & van Pijkeren, J. P. CRISPR-Cas9-assisted recombineering in Author contributions
Lactobacillus reuteri. Nucleic Acids Res 42, e131 (2014). R.B. and P.H. wrote the manuscript.
129. Selle, K. & Barrangou, R. CRISPR-based technologies and the future of food
science. J. Food Sci. 80, R2367–R2372 (2015).
130. Mougiakos, I., Bosma, E. F., de Vos, W. M., van Kranenburg, R. & van der Additional information
Oost, J. Next generation prokaryotic engineering: the CRISPR-Cas toolkit. Reprints and permissions information is available at www.nature.com/reprints.
Trends Biotechnol. 34, 575–587 (2016). Correspondence should be addressed to R.B.
131. Sontheimer, E. J. & Barrangou, R. The bacterial origins of the CRISPR How to cite this article: Barrangou, R. & Horvath, P. A decade of discovery: CRISPR
genome-editing revolution. Hum. Gene Ther. 26, 413–424 (2015). functions and applications. Nat. Microbiol. 2, 17092 (2017).
132. Lander, E. S. The heroes of CRISPR. Cell 164, 18–28 (2016).
Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in
133. Baltimore, D. et al. Biotechnology. A prudent path forward for genomic
published maps and institutional affiliations.
engineering and germline gene modification. Science 348, 36–38 (2015).
134. Zetsche, B. et al. Cpf1 is a Single RNA-guided endonuclease of a class 2
CRISPR-Cas system. Cell 163, 759–771 (2015). Competing interests
135. Shmakov, S. et al. Discovery and functional characterization of diverse class 2 R.B. and P.H. are co-inventors on several patents related to CRISPR–Cas systems and
CRISPR-Cas systems. Mol. Cell 60, 385–397 (2015). their various uses. R.B. is a co-founder and SAB member of Intellia Therapeutics and
136. Burstein, D. et al. New CRISPR-Cas systems from uncultivated microbes. Locus Biosciences, and a shareholder of Caribou Biosciences and DuPont. P.H. is an
Nature 542, 237–241 (2017). employee of DuPont.

NATURE MICROBIOLOGY 2, 17092 (2017) | DOI: 10.1038/nmicrobiol.2017.92 | www.nature.com/naturemicrobiology 9


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.

You might also like