You are on page 1of 16

Molecular Cell

Review

Methods for Optimizing CRISPR-Cas9


Genome Editing Specificity
Josh Tycko,1 Vic E. Myer,1 and Patrick D. Hsu1,2,*
1Editas Medicine, 300 Third Street, Cambridge, MA 02142, USA
2Salk Institute for Biological Studies, 10010 North Torrey Pines Road, La Jolla, CA 92037, USA
*Correspondence: patrick@salk.edu
http://dx.doi.org/10.1016/j.molcel.2016.07.004

Advances in the development of delivery, repair, and specificity strategies for the CRISPR-Cas9 genome en-
gineering toolbox are helping researchers understand gene function with unprecedented precision and
sensitivity. CRISPR-Cas9 also holds enormous therapeutic potential for the treatment of genetic disorders
by directly correcting disease-causing mutations. Although the Cas9 protein has been shown to bind and
cleave DNA at off-target sites, the field of Cas9 specificity is rapidly progressing, with marked improvements
in guide RNA selection, protein and guide engineering, novel enzymes, and off-target detection methods. We
review important challenges and breakthroughs in the field as a comprehensive practical guide to interested
users of genome editing technologies, highlighting key tools and strategies for optimizing specificity. The
genome editing community should now strive to standardize such methods for measuring and reporting
off-target activity, while keeping in mind that the goal for specificity should be continued improvement
and vigilance.

The clustered regularly interspaced short palindromic repeats cations. Initially, bioinformatic solutions for guide RNA selection
(CRISPR)-Cas9 genome engineering technology has been es- recommended guide sequences that recognized cognate
tablished as a powerful molecular tool for numerous areas of bio- genomic DNA targets with minimal sequence homology across
logical study in which it is useful to target or modify specific DNA that given genome. Although such approaches proved reasonably
sequences (Hsu et al., 2014; Sander and Joung, 2014; Cox et al., effective, subsequent systematic profiles of Cas9 off-target activ-
2015; Wright et al., 2016). Originally derived from bacterial adap- ity both in vitro (Pattanayak et al., 2013) and in vivo (Fu et al., 2013;
tive immune systems known as CRISPR-Cas (Mojica et al., 2000, Hsu et al., 2013; Mali et al., 2013a) led to the first set of data-driven
2005; Jansen et al., 2002; Pourcel et al., 2005; Barrangou et al., guidelines for target site selection. Given a particular locus and
2007), the Cas9 nuclease can localize and cleave a target DNA species of interest, computational models and associated web
via a guide RNA (Brouns et al., 2008; Deltcheva et al., 2011; Ga- tools typically rank guide RNAs by predicted specificity and sug-
siunas et al., 2012; Jinek et al., 2012). The Cas9 nuclease has gest likely off-target sites that can be checked for undesired muta-
been engineered to edit target DNA at desired loci in eukaryotic genesis via targeted sequencing or enzymatic assays (see Biased
cells (Cong et al., 2013; Mali et al., 2013c). However, in the Off-Target Detection). However, the appropriate number of poten-
context of large eukaryotic genomes, Cas9 is known to bind tial off-target sites to experimentally assay remains unclear, as the
and cleave at off-target sites (Cong et al., 2013; Fu et al., 2013; accuracy of in silico prediction can vary. Thus, improving methods
Hsu et al., 2013) like its genome editing predecessors zinc finger of off-target detection and quantification has become a focus of a
nucleases (Bibikova et al., 2001; Urnov et al., 2005; Miller et al., number of research groups in the CRISPR-Cas9 field.
2007) and transcriptional activator-like effector nucleases (Boch Several protocols for the unbiased detection of genome-wide
et al., 2009; Moscou and Bogdanove, 2009; Christian et al., Cas9 off-target activity in cells have been recently described
2010; Miller et al., 2011). Efforts to measure, understand, and (Frock et al., 2015; Kim et al., 2015; Ran et al., 2015; Tsai et al.,
improve Cas9 specificity have accordingly been of great interest 2015; Wang et al., 2015) and are important complements to
in the field. While prior review articles should be referred to for biased off-target detection methods based on in silico prediction
foundational background and insight into Cas9 genome editing and targeted sequencing (Fu et al., 2013; Hsu et al., 2013; Mali
development (Makarova and Koonin, 2015) and applications et al., 2013a; Pattanayak et al., 2013) (see Unbiased Genome-
(Mali et al., 2013b; Hsu et al., 2014; Sander and Joung, 2014; wide Off-Target Detection). The adoption of these protocols
Porteus, 2016; Wright et al., 2016), in this article we focus on marks an important departure from methods for measuring an
the sub-field of Cas9 specificity by comparing relevant methods inherently biased subset of potential off-target sites. In addition
for off-target improvement and detection, highlighting useful to these protocols, the computational tools used for guide RNA
experimental and computational tools, and emphasizing further design and off-target prediction are expanding to account for a
research directions for continued improvement. deepening understanding of specificity-governing parameters
Recently, several improvements have been devised for the in the cell, including detailed features of the off-target sequence
wild-type Cas9 enzyme derived from Streptococcus pyogenes, and the Cas9 orthologs (see Computational Guide RNA Selection
currently the nuclease most widely used for genome editing appli- and Predictive Models).

Molecular Cell 63, August 4, 2016 ª 2016 Elsevier Inc. 355


Molecular Cell

Review

Furthermore, multiple methods have been developed to gaps, bulges, or alternative protospacer adjacent motif (PAM) se-
improve specificity by using natural or engineered orthologous quences (sequence motifs recognized by Cas9 directly 30 of the
Cas9s (see Cas9 Orthologs and New CRISPR Proteins), engi- matching guide sequence in the target DNA; typically 50 -NGG
neering the Cas9 protein or guide RNA (see Protein Engineering for SpCas9 but also 50 -NAG with lower efficiency) (Table 1).
and Bespoke PAMs and Modified Guides), or modulating the Looking forward, assays relying on next-generation sequencing
kinetics and regulation of the CRISPR components in the cell of off-target sites may become the standard because of their
(see Kinetics and Regulation). Meanwhile, use of nuclease- sensitivity and accuracy, as current methodologies have a
dead Cas9, or dCas9, for targeted binding applications such reported lower limit of detection of approximately 0.1%, or
as transcriptional regulation have particular considerations given 1/1,000 (Kleinstiver et al., 2016a). It is suggested that the more
the distinctions between the Cas9 target search and nucleolytic accessible Surveyor assay should be used only when quick,
mechanisms (see dCas9 Specificity). CRISPR specificity must first-pass cleavage efficiency measurements are needed. The
also be considered in a broader sense to include cell type and Surveyor assay is generally quantified by analyzing the band
temporal and spatial specificity. Last, the standardization of intensity of gel electrophoresis images and is thought to have a
off-target analysis methods and data reporting in research pub- limit of detection of 5% indels (Fu et al., 2013), making it un-
lications would benefit the field by enabling comparison across suitable for most off-target detection scenarios. Finally, targeted
studies and better algorithms for guide RNA design (see Broader sequencing is perhaps best applied to verify activity at off-target
Implications of Specificity and The Need for Standardization). sites discovered by unbiased, genome-wide methods (Kim et al.,
2015; Ran et al., 2015; Tsai et al., 2015; Kleinstiver et al., 2016a;
Biased Off-Target Detection Slaymaker et al., 2016).
To date, the predominant approach for identifying Cas9
nuclease off-target activity has been to (1) computationally pre- Unbiased Genome-wide Off-Target Detection
dict likely off-target sites on the basis of sequence homology and To address limitations associated with in silico predictions of
then (2) assess any potential editing activity by enzymatic assays potential off-target sites, several unbiased methods have been
on the basis of mismatch-sensitive endonucleases (Guschin recently reported for measuring Cas9 off-target activity across
et al., 2010; Ran et al., 2013b), Sanger sequencing, or targeted the genome (Table 1). We and others have recently reviewed
deep sequencing (see The Need for Standardization). These a and compared these deep sequencing-based protocols (Gabriel
priori predictions are critical, as 98% of Streptococcus pyo- et al., 2015; Gori et al., 2015; Koo et al., 2015; Bolukbasi
genes Cas9 (SpCas9) guide RNAs in human exons and pro- et al., 2016), including direct in situ breaks labeling, enrich-
moters have at least one off-target site with three or fewer ment on streptavidin, and next-generation sequencing (BLESS)
mismatches (Bolukbasi et al., 2016), and foundational Cas9 (Ran et al., 2015), high-throughput, genome-wide translocation
specificity studies collectively demonstrated that off-target sites sequencing (HTGTS) (Frock et al., 2015), genome-wide, unbi-
with three or fewer mismatches are significantly more likely to be ased identification of double-strand breaks (DSBs) evaluated
cleaved than more dissimilar sites (Fu et al., 2013; Hsu et al., by sequencing (GUIDE-seq) (Tsai et al., 2015), integration-defi-
2013; Mali et al., 2013a; Pattanayak et al., 2013). Selecting the cient lentiviral vector (IDLV) capture (Wang et al., 2015), and
most unique target site possible is thus a valuable strategy for in vitro nuclease-digested whole-genome sequencing (Dige-
improving specificity, because many gene editing applications nome-seq) (Kim et al., 2015). These unbiased methods collec-
have several possible guide RNAs capable of accomplishing tively demonstrate that sequence homology alone is not fully
the same experimental outcome (i.e., knocking out a gene by predictive of off-target sites. Further, although the PAM tends
introducing indels in an exon of interest). to be the most stringent predicate of target recognition fidelity
Sequence uniqueness is thus an important initial filter for se- relative to the guide sequence, off-target mutagenesis was
lecting guide RNAs with minimal potential off-target sites with also observed at non-canonical PAMs, confirming that biased
high sequence homology. However, as the number of mis- methods may miss important Cas9 off-target sites (Frock et al.,
matches considered relative to the on-target site increases, the 2015; Kim et al., 2015; Ran et al., 2015; Tsai et al., 2015).
total number of potential off-target sites dramatically increases However, depending on the approach used for enriching
as well. For example, a benchmark guide RNA designed to target Cas9-induced DNA cleavage, unbiased methods tend to be
the EMX1 locus (Ran et al., 2015), with no ‘‘off-by-one’’ mis- less sensitive and have a lower throughput than biased targeted
matched off-target sites in the hg38 reference genome, has 10 sequencing, in addition to typically requiring higher sequencing
‘‘off-by-two’’ sites, 69 ‘‘off-by-three’’ sites, and 27,480 ‘‘off-by- coverage and much more complex protocols. As a result, it is
six’’ sites. currently difficult to generate off-target data at a scale required
As a result, even targeted sequencing, likely the most scalable to adequately train a predictive model for specificity. Regardless,
of the aforementioned approaches, is biased toward a small num- these approaches remain valuable for researchers interested in
ber of sites that can reasonably be assayed. However, Cas9- validating individual guide RNA sequences in their experimental
mediated cleavage has been reported at off-target sites with as systems.
many as six mismatches to the guide sequence (Tsai et al., GUIDE-seq and Digenome-seq have reported sensitivity to
2015), demonstrating the risk of missing detectable events by off-target events as low as 0.1%, or 1 in 1,000 events, whereas
setting low ‘‘off-by’’ thresholds for the sake of practical, biased IDLV capture reported a sensitivity of 1%. The BLESS and
off-target screening. Furthermore, most computational off-target HTGTS studies did not explicitly report summary sensitivity met-
prediction tools do not adequately consider off-target sites with rics. It should be noted that all of these sensitivity rates vary

356 Molecular Cell 63, August 4, 2016


Molecular Cell

Review

depending on the number of cells assayed in the experiment, the turn, drive continued improvement of bioinformatics tools for
deep sequencing depth, and the bioinformatics methods used to target site selection.
differentiate true Cas9-induced cleavage events from back-
ground noise. As a result, differences in reported detection Computational Guide RNA Selection and Predictive
sensitivity between competing protocols do not necessarily Models
reflect inherent limitations of the off-target detection method Many such online tools and applications have been developed
but also a need for technical optimization. for the in silico design of guide RNAs and prediction of their
Beyond sensitivity, interested users should also compare off-target sites (Table 2). Some, such as Cas-OFFinder, simply
the protocols’ applicability in their experimental context (Table identify off-target sites via sequence similarity and rank them
1). For example, GUIDE-seq requires optimized delivery of a on the basis of relative orthogonality, while others use a speci-
defined, chemically modified, double-stranded oligonucleotide ficity score calculation we have previously described (Hsu
(dsODN), making it potentially toxic to certain primary cell et al., 2013). This formula scores guides on the basis of a
types and difficult to apply in vivo on tissue samples. How- weighted sum of the mismatches in their off-target sites, with
ever, the dsODN-specific PCR enriches for DSBs and thus weights experimentally determined to reflect the effect of the
should enable appropriate sequencing coverage with one mismatch position on cleavage efficiency.
sequencing run of reasonable depth. Although BLESS can Other studies have similarly established that mismatches at
be applied to both ex vivo and in vivo samples after Cas9 the 50 end of the spacer are more tolerated than PAM-proximal
editing alone, it requires the biochemical ligation of adaptor mismatches (Cong et al., 2013; Fu et al., 2013; Pattanayak
sequences to free, unrepaired DSB ends. As a result, its sensi- et al., 2013). This ‘‘seed sequence’’ effect has been corroborated
tivity will be affected by the time point after Cas9 editing and by biochemical experiments demonstrating that Cas9 unwinds
DNA repair kinetics. Digenome-seq, on the other hand, can and binds DNA in a directional manner moving away, or 30 to
immediately be applied in any Cas9-editable cell type, yet it 50 , relative to the PAM (Sternberg et al., 2014), and that proper
is an in vitro method that removes genomic structural context. matching of the 50 end of the guide sequence is important for
It originally required expensive whole-genome sequencing proper nuclease domain positioning and successful Cas9 cleav-
(WGS) at high coverage for four sets of genomic DNA per age (Jiang et al., 2015). Beyond weighting for PAM-proximal mis-
assay. An improved version of Digenome-seq lowers the matches, the specificity score also reflects the finding that
cost by multiplexing guide RNAs and reducing the WGS consecutive mismatches attenuate Cas9 activity more than an
requirement from four genomes to one (Kim et al., 2016b). equivalent number of interspaced mismatches. Importantly, as
This improved protocol was run on ten guides that had these target site selection guidelines were derived from studies
been analyzed by GUIDE-seq, and the results were compared on SpCas9, their general applicability to less commonly used
and validated with targeted sequencing of transfected cells. Cas9 orthologs remains unclear.
Although Digenome-seq generally discovered 70 more off- Overall, these bioinformatic tools generally perform well in
target sites per guide initially, the number of validated off- ranking guide RNAs such that, for a given locus, users can easily
target sites was similar for both methods. select guide sequences that are optimized for high genomic
Importantly, as these unbiased methods require manipulation specificity. However, as our mechanistic and empirical under-
of the genome, they will each introduce different and subtle standing of Cas9 activity has progressed, several caveats as
artifacts to the set of identified off-target sites and likely will well as avenues for the improvement of in silico prediction
not report the exact same sites with identical frequencies. The have emerged.
techniques have been compared for some ‘‘benchmark’’ guide For example, there is evidence suggesting off-target sites with
RNAs, but a comprehensive head-to-head comparison is lacking small insertions or deletions in the targeted sequence or alterna-
at this time and may be needed (Gori et al., 2015). The improved tive PAMs can indeed be cleaved by Cas9, albeit with low fre-
Digenome-seq protocol can be rapidly transferred to new quency (Hsu et al., 2013; Jiang et al., 2013; Kleinstiver et al.,
laboratories, as it consists of in vitro Cas9 cleavage and WGS. 2015b; Ran et al., 2015). Cas9 specificity is similarly known to
BLESS and GUIDE-seq, as currently configured, may be the vary because of other variables such as cell type, species, deliv-
most accessible protocols, as they enrich for DSBs and require ery modality, and dosage (see Kinetics and Regulation). Although
fewer sequencing reads per sample. Future generations of these combining existing Cas9 binding and cleavage data sets with
methods should collectively look to improve accuracy and whole-genome chromatin state information has been suggested
throughput as well as reduce the input cell number required. to provide more accurate off-target prediction (Singh et al., 2015),
In sum, these biased and unbiased methods can and perhaps generally such data have been collected for a limited number of
should be applied in a complementary manner. For example, off- guides. It is difficult to integrate all of these observations in a sys-
target sites can be discovered in vitro (e.g., Digenome-seq), in tematic, data-driven way, so these parameters are not taken into
easily cultured cell types (e.g., HTGTS, GUIDE-seq), or even account by the majority of available off-target prediction tools.
in vivo (e.g., BLESS) prior to validation in vivo or in cell lines of in- To realize the full potential of in silico prediction, comprehen-
terest via targeted sequencing. Although combinations of biased sive data sets addressing such limitations must be generated.
and unbiased methods have been applied to a small set of guide Large-scale guide RNA efficiency data sets have been used to
RNAs, larger data sets will be required to determine how well create increasingly accurate computational models that predict
in vitro off-target activity corresponds with Cas9 targeting in vivo the knockout efficiency of new guide RNAs (Doench et al.,
and across different cell types. These types of data sets will, in 2014; Chari et al., 2015; Xu et al., 2015). Most recently, a pooled

Molecular Cell 63, August 4, 2016 357


Molecular Cell

Review

Table 1. Improvements to CRISPR-Cas9 Specificity


Improvement Description Advantage Disadvantage
Measurement
Targeted deep sequencing (Ran targeted amplicon NGS of more quantitative, sensitive, and biased toward a subset of off-
et al., 2013b) putative or known off-target scalable than alternative assays target sites that can reasonably
sites, followed by computational such as Surveyor be assessed; best used as
analysis to quantify the complement to unbiased,
proportion of reads with indels genome-wide off-target
near the PAM site detection methods
GUIDE-seq (Tsai et al., 2015) captures DSBs with a dsODN, straightforward wet-lab protocol requires efficient delivery of the
which is then used as a priming with computational pipelines dsODN, which may be toxic to
site for sequencing available online for data some cell types at some doses,
processing and has not been demonstrated
for in vivo models
BLESS (Ran et al., 2015; biochemical ligation of NGS no exogenous bait is introduced sensitive to time of cell fixation;
Slaymaker et al., 2016) sequencing adapters to to cells; can be applied to tissue current configuration requires
exposed gDNA ends; samples from in vivo models relatively large number of cells
computational filtering
separates Cas9-mediated DSBs
(aligned near PAM) from
naturally occurring DSBs (more
randomly distributed)
Digenome-seq (Kim et al., 2015; cell-free gDNA samples are can be applied to any cell type as WGS may be costly; must be
Kim et al., 2016b) digested in vitro by Cas9 RNP digestion performed on paired with one of the other
with multiplexed guide RNAs extracted gDNA; more sensitive methods to validate the sites
before WGS; cut sites are than GUIDE-seq in a head-to- discovered in vitro are truly
identified on the basis of read head mutagenized in the cell
alignment
Computational Guide Selection
Specificity score (Hsu et al., an experimentally derived a continuous variable to score is derived from SpCas9
2013) and CFD (Doench et al., function to score the likelihood of distinguish guide RNAs during data with 20-mer guides, so it
2016) an off-target site being edited, design and to rank off-target may not generalize to other
on the basis of the position and sites for targeted sequencing contexts
nature of its mismatches follow-up
WGS of reference genome (Yang WGS of the relevant cell line, identify new off-target sites remains costly; custom
et al., 2014) animal model, patient, etc. created by genetic variation, reference genomes only
(Box 1) which are not present in the available with some guide
reference genomes design tools (Table 2)
(i.e., hg38)
Protein Engineering
Single/paired nickases (Cong a point mutation at the active site nicks are repaired with higher less efficient on-target editing
et al., 2013; Mali et al., 2013a; of one of the Cas9 nuclease fidelity than DSBs, so off-target with some guide RNAs
Ran et al., 2013a) domains yields a targeted edits are less frequent; nickases
nickase; paired nickases can also mediate efficient
targeting complementary homology directed repair
strands create a deletion
SpCas9 PAM Variant D1135E a single point mutation increases significant decrease in editing at on-target efficiency may be
(Kleinstiver et al., 2015b) specificity for the 50 -NGG PAM 50 -NAG and 50 -NGA PAMs; affected; it was comparable
improved genome-wide with wtCas9 for six guides by
specificity for several guides as T7E1 but somewhat lower for
assessed by GUIDE-seq three guides by deep
sequencing
eSpCas9 (Slaymaker et al., three mutations within the off-target editing was nearly on-target efficiency may be
2016) nt-groove weaken Cas9’s non- entirely avoided, as assessed by affected, although it was
target DNA strand stabilization both BLESS and targeted deep comparable with wtCas9 for
and therefore increase sequencing most of the guide RNAs reported
stringency of guide RNA-DNA
complementation for nuclease
activation
(Continued on next page)

358 Molecular Cell 63, August 4, 2016


Molecular Cell

Review

Table 1. Continued
Improvement Description Advantage Disadvantage
SpCas9-HF (Kleinstiver et al., four mutations weaken Cas9 off-target editing was nearly on-target efficiency may be
2016a) binding to the target DNA strand entirely avoided, as assessed by affected, although it was
and therefore increase both GUIDE-seq and targeted comparable with wtCas9 for
stringency of guide RNA-DNA deep sequencing most of the guide RNAs reported
complementation for nuclease
activation
RNA Modifications
tru-guides (Fu et al., 2014b) tru-guides have spacers of simple to implement; can less efficient on-target editing
17–18 nt as opposed to the decrease activity at genomic with some guide RNAs; some
canonical 20 nt; the shorter off-target sites guides may have more off-target
region of RNA:DNA sites with high-sequence
complementarity allows fewer homology
mismatches
50 GGX20 guides (Cho et al., two additional mismatched this simple modification can be this modification is compatible
2014) guanines are added to the 50 end rapidly assessed; specificity only with certain guide RNAs;
of the guide RNA; these PAM- ratios at off-target sites generally on-target efficiency is
distal mismatches may adjust were reported to improve by sometimes severely decreased
binding energetics such that 10- to 100-fold or more
Cas9 more stringently requires
complementarity
Delivery
SaCas9 (Ran et al., 2015) the Staphylococcus aureus Cas9 smaller size allows AAV the ortholog is less
ortholog is 3.2 kb, uses a 20–23 packaging with one or two characterized than SpCas9
nt guide RNA, and recognizes a sgRNAs in an all-in-one vector; and may be less efficient
50 -NNGRRT PAM specificity improvements were on-target in some contexts
observed with some guides by
BLESS, GUIDE-seq, and
targeted sequencing
RNP (Cho et al., 2013b; purified Cas9 protein is Burst-like Cas9 expression best used ex vivo
Kim et al., 2014; Lin et al., 2014; complexed with guide RNA kinetics mediates efficient on-
Zuris et al., 2015) before delivery to cells by lipid target editing without long-term
transfection or electroporation risk of accumulating off-target
edits
gDNA, genomic DNA; NGS, next-generation sequencing; tru-guides, truncated guides.

screen analysis of 9,914 guide RNA variants covering all single Last, the online tools compared in Table 2 will be of practical
mismatch and single RNA indels for 65 DNA targets was used utility for most users, but bioinformaticians may benefit from
to derive a cutting frequency determination (CFD) score capable the performance and flexibility of offline tools, such as CRISP-
of predicting the likelihood of off-target activity (Doench et al., Rseek (Zhu et al., 2014). A comprehensive comparison of these
2016). computational tools can be found with the CRISPR Software
Overall, these data underscore the complexity of predicting Matchmaker tool (MacPherson, 2015).
Cas9 specificity and the importance of considering both position
and mismatch identity in metrics like the CFD. For example, 256 Cas9 Orthologs and New CRISPR Proteins
of 289 (89%) rG:dT mismatched guides retained high activity, Recently, a broader array of Cas9 ortholog proteins have been
whereas only 37% of rC:dC mismatches were active. Looking engineered as genome editing tools, such as the Staphylo-
just at position 16 in the guide sequence, all 46 purine-purine coccus aureus Cas9 (SaCas9), which is more than 1 kb smaller
mismatches tested attenuated Cas9 activity, whereas 97% of than the more commonly used SpCas9 and thus easily pack-
rN:dT mismatches were tolerated. Furthermore, in a validation aged into AAV paired with guide RNA expression cassettes for
study, modest correlation was reported between the CFD and in vivo genome editing (Ran et al., 2015; Nishimasu et al.,
observed off-target activity at 60 sites with one or two mis- 2015; Friedland et al., 2015). Interestingly, SaCas9 demonstrates
matches. Generally, well-designed guide RNAs avoid off-by- highly efficient nuclease activity with target sequences of
one and off-by-two mismatch sites, so these models should be 21–24 nt compared with the 20 nt targets used with SpCas9,
improved to predict activity at off-by-three and off-by-four while using a 50 -NNGRRT PAM as opposed to the 50 -NGG
mismatch sites. Future refinements and insights will likely be of SpCas9. SaCas9 also demonstrates substantial cleavage
derived from large training data sets that consider off-target sites efficiency at 50 -NNGRR PAM sites. It remains to be determined
with more than one mismatch or indel. how a longer target sequence and longer PAM preference

Molecular Cell 63, August 4, 2016 359


Molecular Cell

Review

Table 2. Computational Tools for Designing Guide RNAs


Single/ Cas9 Custom Lists Considers
Genomes Paired Orthologs Guide Off- Off-Targets
Name Website Specificity Score Supported Design Supported Length Targets with Indels
Benchling http://benchling.com Hsu et al. (2013) and CFD 22 both 5 yes yes no
(Doench et al., 2016)
BLAST (on pre- http://blast.ncbi.nlm.nih.gov/ sequence similarity R1,000 NA NA yes yes yes
designed guide) Blast.cgi
Cas-OFFinder http://www.rgenome.net/ no 34 single 4 yes yes yes
cas-offinder/
MIT CRISPR http://CRISPR.mit.edu Hsu et al. (2013) 16 both 1 no yes no
Design Tool
Deskgen http://deskgen.com Hsu et al. (2013) 5 both 4 yes yes no
DNA 2.0 CRISPR https://www.dna20.com/ relative ranking based 5 both 1 no no no
gRNA Design Tool eCommerce/cas9/input on orthogonality
E-CRISP http://www.e-crisp.org/ relative ranking based 34 both 1 yes yes no
E-CRISP/designcrispr.html on orthogonality
EuPaGDT http://grna.ctegd.uga.edu relative ranking based 28 single custom yes yes no
on orthogonality PAM
GenScript gRNA http://www.genscript.com/ no 2 single 1 yes no no
Design Tool gRNA-design-tool.html
ZiFiT http://zifit.partners.org/ZiFiT/ orthogonality of off- 9 both 1 yes yes no
Disclaimer.aspx targets
Broad GPP http://www.broadinstitute.org/ CFD (Doench et al., 2 single 1 no no yes
Portal rnai/public/analysis-tools/ 2016)
sgrna-design
CROP-IT http://cheetah.bioch.virginia. Singh et al. (2015) 2 both 1 no yes no
edu/AdliLab/CROP-IT/

affects the genome-wide off-target activity of SaCas9, although genome databases for type II CRISPR loci, more than 56 bacte-
initial results with the unbiased BLESS and GUIDE-seq protocols rial species have been identified with predicted novel trans-acti-
appear promising (Friedland et al., 2015; Ran et al., 2015). Given vating CRISPR RNAs (tracrRNAs) (Chylinski et al., 2013).
identical guides targeting sites with nested 50 -(NGG)RRT PAMs, Furthermore, Cas9 proteins phylogenetically clustered within
editing with SaCas9 as opposed to SpCas9 resulted in lower in- the same subgroup can interchangeably use their crRNA-
del frequencies at fewer off-target sites (Ran et al., 2015). tracrRNA (dual-RNA) duplexes to cleave DNA, suggesting that
Cas9s derived from Neisseria meningitidis (NmCas9) and Cas9-guide RNA pairings could be optimized by mining the
Streptococcus thermophilus (StCas9) CRISPR1 and CRISPR3 natural diversity or engineering improved guide architectures
have also been engineered for specific genome editing in (Fonfara et al., 2014).
mammalian cells (Mali et al., 2013a; Müller et al., 2016). Similar Furthermore, the large natural diversity of bacterial CRISPR
to SaCas9, NmCas9 requires a more complex 50 -NNNNGATT systems extends beyond type II systems featuring Cas9 as the
PAM and uses a longer crRNA spacer (21–24 nt) than SpCas9 sole effector nuclease. Recently, a newly discovered CRISPR
(Hou et al., 2013; Mali et al., 2013a; Lee et al., 2016). Impor- protein named Cpf1 has also been used for human genome edit-
tantly, although the longer PAM limits the targeting range, it ing (Zetsche et al., 2015). Also targeted by a single guide RNA
also limits the number of off-target sites for any given guide. (sgRNA), Cpf1 recognizes a 50 -TTN PAM upstream (rather than
For example, NGG motifs occur roughly every 8 bp, whereas downstream) of the target site and creates a 5 nt staggered over-
NNNNGATT is found every 128 bp (Lee et al., 2016). A system- hang DSB distal to the PAM (Fonfara et al., 2016; Yamano et al.,
atic study of NmCas9 mismatch and indel tolerance found an 2016). It is important to note that eight Cpf1 orthologs were
overall improvement over SpCas9 and a promising absence screened in human cells, and only two mediated efficient cleav-
of any off-target editing at all sites with three or fewer mis- age. Cpf1 genome-wide specificity was recently assayed with
matches (Lee et al., 2016). However, the on-target efficiency Digenome-seq (Kim et al., 2016a) and GUIDE-seq (Kleinstiver
is consistently below that of SpCas9, possibly because of et al., 2016b) and can be comparable with or greater than
weaker DNA unwinding activity (Ma et al., 2015). The two SpCas9 for many guide RNAs. Other putative CRISPR effector
StCas9 orthologs similarly might be more specific than the proteins, known as c2c1, c2c2, and c2c3, have also been bio-
wild-type SpCas9, but they also suffer from weaker efficiency informatically mined from metagenomic data sets (Shmakov
(Müller et al., 2016). et al., 2015). C2c2 has since been characterized as a program-
Overall, Cas9 orthologs with improved specificity profiles can mable RNA-guided RNA-targeting nuclease in bacterial cells
likely be discovered and engineered. By screening available (Abudayyeh et al., 2016). Interestingly, once c2c2 binds its

360 Molecular Cell 63, August 4, 2016


Molecular Cell

Review

complementary target RNA, it is active as a non-specific RNase de-coupling of efficiency and specificity without discretely modi-
and significantly impedes cell growth. It remains largely unclear fying the mechanism of action (as accomplished by the cooper-
how these new CRISPR proteins’ different characteristics (i.e., ativity strategies).
guide RNAs, PAMs, nuclease domains) will affect their speci- A similar strategy by a different research group resulted in
ficity, but their discovery reflects great potential for genome en- a ‘‘high-fidelity’’ SpCas9-HF (Kleinstiver et al., 2016a). In
gineering beyond SpCas9. contrast with eSpCas9, SpCas9-HF contains four alanine
substitutions at residues that interact with the phosphate
Protein Engineering and Bespoke PAMs backbone of the targeted DNA strand. Although on-target ef-
Meanwhile, biological engineers have rapidly optimized and ficiency appeared to be slightly more variable than eSpCas9
successfully matured SpCas9 as a tool, with increasing preci- for the guides tested, it remained comparable (>70%) with
sion as understanding of its molecular structure and mecha- wild-type efficiency for 86% of 37 guide RNAs. Altogether,
nism has deepened (Table 1). The first generation of improve- these engineered Cas9s greatly improve specificity without
ments leveraged knowledge of the two nuclease domains of much sacrifice in efficiency. Interestingly, these improved var-
Cas9, initially identified on the basis of sequence homology iants present a new challenge to improve the limit of detection
(Sapranauskas et al., 2011). For example, a single point muta- of biased and unbiased assays beyond 0.1%, in order to cap-
tion (D10A) can inactivate the RuvC domain and change the ture remaining rare off-target events that may be important in
SpCas9 nuclease into a nickase capable of cleaving only the clinical applications.
DNA strand complementary to the guide RNA (Cong et al., Another successful approach has been to alter the PAM
2013). Using paired guide RNAs to create two appropriately recognition sequence of Cas9 itself. Because the PAM is
offset nicks results in efficient indel formation while improving perhaps the most stringent determinant of Cas9 specificity,
specificity by as much as 1,500-fold, likely because single Cas9s with alternative PAMs could not only increase the number
DNA nicks are repaired with high fidelity compared with DSBs of targetable sites across a genome of interest but also poten-
(Mali et al., 2013a; Ran et al., 2013a; Friedland et al., 2015). tially demonstrate improved genome-wide specificity by
Paired nickases have been applied to efficiently generate requiring a longer PAM or a PAM with less abundance in the
mouse models with no detectable off-target edits, whereas target genome. One strategy is to swap the putative PAM-inter-
wild-type Cas9 can generate off-target edits that are inherited acting domain (PID) of the protein for the PID from an ortholog
by founders’ progeny (Shen et al., 2014). Beyond nickases, a recognizing a different PAM. This has been previously demon-
second point mutation in the HNH catalytic domain creates strated with SpCas9 and Streptococcus thermophilus CRISPR3
a fully nuclease-inactive ‘‘dCas9’’ (Nishimasu et al., 2014). By Cas9 (St3Cas9), which share 60% sequence identity and can
fusing dCas9 to the FokI nuclease domain and again expressing swap dual guide RNA sequences while retaining function (Fon-
paired guide RNAs, one can create a dimerization-dependent fara et al., 2014; Nishimasu et al., 2014).
system that also improves specificity (Guilinger et al., 2014; Similarly powerful is the leveraging of directed evolution to alter
Tsai et al., 2014). the PAM specificity, as shown recently for SpCas9 (Kleinstiver
Another recent protein engineering approach, known as et al., 2015b) and SaCas9 (Kleinstiver et al., 2015a). Notably,
SpCas9MT-pDBD fusions, involves linking a mutated Cas9 with just four mutations resulted in a modified ‘‘VRER’’ SpCas9 with
attenuated PAM-binding (SpCas9MT) to a programmable DNA- high specificity for NCGC PAMs, which are roughly 23 times
binding domain (pDBD) such as a ZF or TALE that targets nearby less abundant in the human genome than the canonical NGG
genomic DNA (Bolukbasi et al., 2015). SpCas9MT-pDBD pro- PAM for standard 50 -G-N19 guide RNAs. Directed evolution also
vided specificity improvements up to 150-fold, likely by yielded a single point mutation (D1135E) SpCas9 that increases
leveraging cooperativity between two independent DNA-binding specificity for NGG PAMs over the alternative NAG PAM (Hsu
events to improve specificity, much like the aforementioned et al., 2013; Jiang et al., 2013; Kleinstiver et al., 2015b). In the
nickase and FokI strategies. case of SaCas9, directed evolution was applied to modify the
However, these strategies have the overall disadvantage of PAM from NNGRRT to NNNRRT, in order to increase the target-
requiring additional components and larger transgenes. ing range by 2- to 4-fold (Kleinstiver et al., 2015a).
Recently, a crystal structure-based, rational engineering strat- One advantage of this unbiased directed evolution approach
egy generated SpCas9 and SaCas9 nucleases with ‘‘enhanced is that it does not require structural data and can be rapidly
specificity’’ (eSpCas9 and eSaCas9, respectively). These applied to diverse orthologs. It could prove important for thera-
Cas9s differ by only three and four codon substitutions, respec- peutic targets requiring a very specific edit and lacking nearby
tively, within the non-targeted DNA strand groove (nt-groove) of canonical PAMs. Finally, bespoke PAMs are likely compatible
Cas9 (Slaymaker et al., 2016). These mutations are thought to with the enhanced specificity and high fidelity mutations. Chi-
weaken non-target strand binding by the protein, encouraging meras, directed evolution, and other protein engineering ap-
DNA rehybridization in competition with guide RNA invasion of proaches may also be fruitful for generating Cas9s that recog-
the target DNA strand (Figure 1). With increased stringency of nize AT-rich PAMs, which could be useful for targeting intronic
RNA-DNA matching, these Cas9s demonstrate dramatic reduc- regions or AT-rich genomes.
tion in the number of genome-wide off-target sites detectable by
BLESS or follow-up targeted sequencing. Moreover, eSpCas9’s dCas9 Specificity
on-target efficiency appeared to be comparable with wild-type Aside from DNA cleavage, a large collection of Cas9 applications
SpCas9 for the 24 guide RNAs tested, suggesting a surprising use catalytically inactivated Cas9 (dCas9), which can be fused to

Molecular Cell 63, August 4, 2016 361


Molecular Cell

Review

A B

Figure 1. Model of Cas9 DNA Targeting and Cleavage


(A) Cas9 in complex with guide RNA samples the genome for PAM matches by diffusion and undersamples heterochromatin regions under certain conditions
(Knight et al., 2015).
(B) Upon PAM recognition and double-stranded DNA binding, the DNA is directionally unwound as the guide RNA interrogates the target strand for comple-
mentary sequences.
(C) The non-target single-stranded DNA (ssDNA) strand is stabilized in a non-sequence-specific manner in Cas9’s nt-groove. Incomplete complementarity or a
short 15 nt PAM-proximal region of complementarity is sufficient for Cas9 binding, as demonstrated by dCas9 or ‘‘dead guides’’ with wtCas9, respectively. Near-
complete complementarity of >17 nt of RNA:DNA heteroduplex is required for nuclease activation.
(D) After this RNA strand invasion, the HNH domain undergoes a conformational change and communicates through a linker to the RuvC domain. The HNH and
RuvC catalytic domains simultaneously cleave the target and non-target ssDNA strand, respectively. Cas9 may remain bound to the cut site for an extended
period before returning to the pre-target state (Richardson et al., 2016).
The model shown is adapted from several sources (Anders et al., 2014; Nishimasu et al., 2014; Sternberg et al., 2014, 2015; Jiang et al., 2016; Slaymaker et al.,
2016).

various effector domains for applications ranging from program- for guide RNA-target DNA complementarity until the nuclease
mable chromosomal labeling to targeted transcriptional and achieves an active state upon full guide RNA-target hybridization
epigenetic control (Hsu et al., 2014). However, the specificity (Jiang et al., 2013, 2015).
profile of dCas9 binding versus wild-type (wtCas9) cleavage ap- In accordance with this model, dCas9 immunoprecipitation
plications needs to be carefully considered. Recent work study- studies demonstrate a high promiscuity of Cas9 binding that cor-
ing the target search mechanisms (Figure 1) of the active Cas9 relates poorly with Cas9 cleavage (Kuscu et al., 2014; Wu et al.,
enzyme indicates that Cas9 initially scans the genome for PAM 2014; Ran et al., 2015): across nearly 300 off-target binding loci
sites (Sternberg et al., 2014; Jiang et al. 2015), resulting in a tran- identified via dCas9 chromatin immunoprecipitation sequencing
sient binding state stabilized by a 5 bp seed sequence of the (ChIP-seq), only 1 of these sites was mutated slightly above
guide RNA (Wu et al., 2014). Additional bases are interrogated background level by wtCas9 (Wu et al., 2014). In a GUIDE-seq

362 Molecular Cell 63, August 4, 2016


Molecular Cell

Review

genome-wide study of DNA cleavage, only 2% of off-target sites tigation into guide RNA features will provide a basis for the opti-
intersected with a set of 585 dCas9 ChIP-seq off-target binding mization of guide RNAs for specificity, functionalization, and
sites (Tsai et al., 2015). Moreover, dCas9 ChIP-seq identified other applications.
sites with about seven mismatches on average, whereas cleav- The length of the guide sequence itself is also an important
age sites harbored about three mismatches on average (Tsai mediator of specificity. Although extending the 20 nt guide
et al., 2015). Although a useful unbiased genome-wide assay sequence of SpCas9 to 30 nt appeared to result in processing
for specificity of dCas9 applications, one should not consider back to the natural 20 nt length (Ran et al., 2013a), shortening
dCas9 ChIP-seq as a proxy for genome-wide detection of the guide sequence proved to be a successful strategy. Spacer
‘‘true’’ cleavage events. sequences of 17 or 18 nt have been reported to mediate more
Just as for its nuclease counterpart, a combination of biased precise genome editing, potentially by increasing the sensitivity
and unbiased approaches provides the most reliable assess- of Cas9 binding to mismatches within the shorter complemen-
ment of dCas9 off-target activity. One method is a mismatch re- tary sequence (Fu et al., 2014b). Although truncated guides
porter assay in which a library of mismatched target sites is can eliminate activity at many off-target sites, they can also
cloned upstream of a promoter driving expression of barcoded create new off-target sites because of their shorter length (Fu
RNAs (Mali et al., 2013a). RNA sequencing (RNA-seq) is then et al., 2014b; Tsai et al., 2015; Wyvekens et al., 2015; Slaymaker
used to determine the extent to which that mismatched target et al., 2016). Interestingly, truncated guides are not compatible
site enabled dCas9-mediated activation or repression of the pro- with the ‘‘enhanced specificity’’ nt-groove mutants or SpCas9-
moter. This method revealed that dCas9-effector fusions are HF, as the combination results in very low on-target efficiency
largely insensitive to point mutations and can repress or activate (Kleinstiver et al., 2016a; Slaymaker et al., 2016). It remains to
transcription while tolerating up to three mismatches, especially be seen if this simple guide design modification could improve
near the 50 end of the guide RNA. That mismatch tolerance may the specificity of other Cas9 orthologs while maintaining editing
be less than that measured by ChIP-seq because of the longer efficiency, which also varies by guide length (Friedland et al.,
duration of promoter binding required to make the transcriptional 2015; Ran et al., 2015).
regulatory function detectable. Notably, shortening the guide further to 14 or 15 nt creates a
Whole-transcriptome RNA-seq also works as an unbiased ‘‘dead guide’’ that can mediate efficient wild-type Cas9 binding
method for dCas9-effector off-target analysis (Perez-Pinera without nuclease activation, enabling multiplexed applications
et al., 2013; Konermann et al., 2015; Polstein et al., 2015). Gener- of DNA cleavage and targeted effector activity (Dahlman et al.,
ally, this assay confirms the highly specific upregulation of the 2015; Kiani et al., 2015). This engineered guide approach
genes targeted by dCas9 activators. This specificity is unsurpris- benefited from insight into the mechanism of Cas9 cleavage,
ing given that dCas9 activators are efficient only within a limited which has been shown to require near complete strand invasion
window of the transcription start site: binding events >200 nt up- and complementarity (Figure 1). Whole-transcriptome RNA-seq
stream of a start site will have little effect on transcription (Gilbert revealed highly similar specificity profiles for both 15 and 20 nt
et al., 2014; Konermann et al., 2015). Notably, one RNA-seq guides applied for transcriptional activation (Dahlman et al.,
study identified a moderate yet significant downregulation of 2015). Although ‘‘dead guides’’ may have more predicted off-
IL32 mRNA, a proinflammatory cytokine, in response to target binding sites, they likely exhibit acceptable specificity
dCas9-VP64 transcriptional activators even without a guide (fewer than three off-target genes significantly activated)
RNA (Perez-Pinera et al., 2013). This result emphasizes the because off-target sites must fall within a limited window of a
importance of considering off-target effects of Cas9 beyond transcription start site in order to perturb transcription.
the first-order risk of binding or cutting activity at a degenerate Another simple modification, known as ‘‘GGX20,’’ adds two
target site, such as unintended biological activity in response additional mismatched guanine nucleotides to the 50 end of the
to the introduced genome editing components (see Broader Im- guide RNA (Cho et al., 2014). Several studies have demonstrated
plications of Specificity). specificity improvements of 10- to 100-fold or greater, although
the on-target efficiency varies from comparable to severely
Modified Guides reduced (Cho et al., 2013a, 2014; Sung et al., 2014; Kim et al.,
Although the aforementioned strategies engineer the Cas9 pro- 2016b). Despite such variability, ‘‘GGX20’’ and truncated guides
tein, specificity improvements may also be achieved by modi- are easily generated and assessed, although they should still be
fying the guide RNA (Table 1). Recent characterization of single compared with standard-length guides on a case-by-case basis
and dual-guide RNA systems for type II CRISPR systems identi- to optimize on-target efficiency.
fied six functional modules within the guide (Briner et al., 2014). Finally, chemical modification of guide RNAs can also
These modules can be mutated, extended, and deleted to impact Cas9 specificity. Synthesis of guide RNAs with 20 -O-
modulate on-target efficiency. For example, extending the methyl (M), 20 -O-methyl 30 phosphorothioate (MS), or 20 -O-
sgRNA duplex region by 5 nt and/or mutating the string of four methyl 30 thioPACE (MSP) at the 50 and 30 ends significantly
T’s shortly downstream of the spacer can improve efficiency improves editing efficiency in human primary T cells and
by changing the RNA’s structure and transcription rate, respec- CD34+ hematopoietic stem and progenitor cells (Hendel et al.,
tively (Chen et al., 2013; Dang et al., 2015). This approach also 2015). However, these RNA modifications increased off-target
improves efficiency for SaCas9 sgRNAs, indicating the general activity at some sites. In a contrasting study, synthetic guide
applicability of guide modifications across multiple Cas9 ortho- RNAs with 10 M-modified 50 bases and interspersed 20 -fluoro
logs (Chen et al., 2016; Tabebordbar et al., 2016). Further inves- modifications toward the 30 end were more stable and specific

Molecular Cell 63, August 4, 2016 363


Molecular Cell

Review

but less efficient on-target (Rahdar et al., 2015). Given these molecule (Davis et al., 2015). However, the on-target efficiencies
mixed early findings, the ability to modulate RNA stability and were also generally less than that of wild-type Cas9, again
mismatch binding kinetics through alternative nucleotide chem- demonstrating the cost associated with some efforts to engineer
istries of guide RNAs warrants further exploration. improved specificity. Small molecules can also be used to
modulate the cellular response to genome editing. For example,
Kinetics and Regulation the rate of homology-directed repair over non-homologous
A critical and underexplored determinant of Cas9 specificity is end-joining has been improved by using a DNA ligase IV inhibitor
the pharmacokinetic profile of the delivered components across (Maruyama et al., 2015) and other molecules with poorly under-
cell and tissue types, especially the form factor of the genome stood mechanisms of action (Yu et al., 2015). Additionally,
editing components (DNA, RNA, or protein) and the delivery synchronizing the cell cycle has been suggested to increase
vehicle (viral or non-viral). the homology-directed repair rate with RNP delivery in human
Lentiviral-mediated integration of Cas9 and guide RNA into cells (Lin et al., 2014).
melanoma cell lines has been reported to achieve saturating in-
dels up to 100% at on-target sites (Shalem et al., 2014). Although Broader Implications of Specificity
both on- and off-target activity generally appeared stable be- Proper consideration of Cas9 genomic specificity should include
tween days 7 and 14, indels at one off-target site with three mis- not only the aggregate number of potential off-target sites for a
matches rose from 40% to 50%. Because off-target levels given guide RNA but also the physiological impact of individual
can increase with sustained exposure to Cas9, controlling or off-target events (both observed and not). Although tumorigen-
limiting the time of Cas9 expression in the cell or nucleus may esis and unwanted editing of oncogenic tumor suppressors is
improve specificity. a common concern, editing events that negatively impact the
As such, the Cas9 delivery modality is an important mediator cell’s viability or functional capabilities also need to be consid-
of specificity (Table 1). For example, cationic lipid-mediated ribo- ered. Moreover, the ability to measure off-target editing differs
nucleoprotein (RNP) delivery improved the specificity of three significantly by tissue. Although blood can be sampled relatively
SpCas9 guides at 11 assessed off-target sites (typically by easily for genomic DNA sequencing, other tissues cannot be
10-fold) compared with plasmid DNA transfection, after normal- so simply biopsied. As a result, linking observed phenotypes
izing for equivalent on-target efficiency (Zuris et al., 2015). In a with Cas9-induced genome modifications could be difficult in
similar study, RNP-Cas9 improved the on-target/off-target ratio certain organs. The ‘‘acceptable’’ thresholds for off-target rates
by approximately 3-fold for two guides tested (Kim et al., 2014). could also differ in terminally differentiated cells, as opposed to
Delivering MS-modified synthetic guide RNAs as RNPs signifi- dividing cells. Certainly, ‘‘acceptable’’ specificity thresholds may
cantly improved on/off-target ratios at all 3 assessed sites not be universal across applications, if they are even determin-
compared with co-delivery with Cas9 plasmid or mRNA (Hendel able, as ultimately empirical demonstration of safety will be of
et al., 2015). This specificity benefit is possibly due to the burst- paramount importance in the clinic.
like kinetics of RNP delivery and associated Cas9 exposure as Furthermore, the specificity of Cas9 nucleases is not neces-
opposed to more stable plasmid-driven expression (Kim et al., sarily confined to genomic specificity, especially for in vivo appli-
2014). cations. For example, tissue and cell-type specificity are also
Cas9 dosage can also affect the kinetics and modulate spec- important considerations when targeting genetic disorders that
ificity, both in vitro (Pattanayak et al., 2013; Fu et al., 2014a) and primarily affect certain organs or cells. However, selective distri-
in cells (Hsu et al., 2013). For example, by plasmid transfection, a bution may be mediated by a careful choice of the delivery
5-fold drop in dose led to a 7-fold increase in the specificity ratio method. For example, different AAV serotypes present different
at the cost of a 2-fold decrease in on-target efficiency, empha- cell-type tropisms and blood clearance rates when administered
sizing the importance of titrating delivery conditions to optimize systemically (Zincarelli et al., 2008). Alternative routes of admin-
specificity for any given application (Hsu et al., 2013). Using the istration can also provide localized or tissue-specific expression,
weaker H1 promoter for guide RNA expression gives a greater as well as immunogenicity benefits (Ge et al., 2001; Limberis and
decrease in off-target activity than on-target activity (Rangana- Wilson, 2006). Whereas gene augmentation therapy approaches
than et al., 2014). Such non-linearity in the on-target/off-target previously demanded viral serotypes that provided sustained
editing rate relationship with Cas9 or guide RNA concentration and high transgene expression (Hastie and Samulski, 2015),
reflects the significant, yet incomplete, coupling of the target- Cas9 applications may be safer with vectors that deliver short-
finding and nuclease activities of the enzyme. Single-molecule term expression in a highly specific cell type.
imaging studies performed with very low Cas9 concentrations Cell-type specificity may also be achieved with transcriptional
revealed undersampling of heterochromatin regions, suggesting regulatory elements. RNA polymerase II tissue-specific and cell
that DNA site accessibility can also play a role in specificity under type-specific promoters (e.g., TBG, Synapsin, GFAP) capable
certain conditions (Knight et al., 2015). of driving protein-coding transcripts such as Cas9 could be
Small molecules can also modulate the kinetics of CRISPR combined with novel guide RNA promoters and expression stra-
editing. For example, a small molecule-dependent Cas9 was tegies beyond the U6 and H1 RNA polymerase III promoters long
developed by genetically inserting an evolved ligand-dependent used in the RNAi field for expressing small interfering RNA
intein, which increased specificity upon induction by 2- to 25- (siRNA) and small hairpin RNA transcripts. A Csy4-dependent
fold at 11 assessed off-target sites for three guide RNAs, while method allows RNA polymerase II promoters to express guide
showing very limited nuclease activity in the absence of the small RNAs (Nissim et al., 2014), permitting simple multiplexing as

364 Molecular Cell 63, August 4, 2016


Molecular Cell

Review

Box 1. Specific for Which Genome?

The simplest way to avoid most off-target effects is to design guide RNAs that have few off-target sites with high sequence ho-
mology (i.e., zero, one, or two mismatches). However, mutations in the genome or the guide RNA can theoretically create off-target
sites with greater sequence homology than expected.

MUTATIONS IN THE GENOME

It can be important to consider the reference genome used with these design tools. The current standard practice is to design
guide RNAs against the species of interest’s standard reference genome (i.e., hg38), yet variation in the particular genome of
the edited tissue, cell line, or strain can be an important determinant of guide specificity. In fact, one group recently observed
an off-target event of unexpectedly high activity (36.7%) when targeting human induced pluripotent stem cells, because of a
SNP that converted a three-mismatch site into a two-mismatch site on one allele (Yang et al., 2014). For CRISPR-Cas9 applications
that require exquisite specificity, WGS and an understanding of patient or genotypic variation may become an important initial step
of guide design.

MUTATIONS IN THE GUIDE RNA

Guide RNAs can be introduced to the cell in several ways with varying fidelity. For example, DNA oligos are prone to small dele-
tions, as the coupling reaction that adds each base is 99% efficient. Digenome-seq reported ‘‘RNA-bulge’’-type off-targets with
guides transcribed from an annealed-oligo template but no such off-target sites with guides transcribed from a clonal plasmid DNA
template (Kim et al., 2016b). Effectively, 1 nt deletions in the oligo pool were leading to ‘‘false positives’’ at off-target sites that
would not be edited with other delivery methods. When using RNP delivery, note that RNA oligo synthesis is less efficient
(98%) and thus more prone to deletions. As an alternative, one can transcribe RNA in vitro from sequence verified plasmid
with T7 polymerase, which has a more acceptable error rate of 1/20,000 (Huang et al., 2000). The error rate of Pol III transcription
is 1 3 107 in yeast and may be similarly low for guide RNAs transcribed from the human U6 promoter (Alic et al., 2007). Thus,
guide RNAs expressed by plasmid and viral delivery likely boast high fidelity, but direct delivery of synthesized oligos should be
examined closely for some guide RNAs.

well as expression restricted to certain cell types. Additionally, for ever improving enzymes and measurement technologies
microRNA binding sites could be added to the 30 end of the with improved lower levels of detection while carefully assessing
Cas9 mRNA to limit expression in certain cells. Reported to be candidate therapies using state of the art functional models and
antigen-presenting cell (APC)-specific, mir-142-3p has been ex- assays will be important to making appropriate pre-clinical deci-
ploited in this manner to effectively limit antibody and cellular re- sions for genome editing therapies.
sponses to transgenes by repressing expression and peptide To rigorously benchmark progress, the genome editing com-
presentation in APCs (Majowicz et al., 2013). munity should strive to standardize the best-performing
Finally, another means of controlling the specificity of Cas9- methods for measuring and reporting off-target activity (Joung,
mediated activity may be through optogenetic control. As previ- 2015). For example, the commonly used Surveyor assay inher-
ously reported with TALEs (Konermann et al., 2013), harnessing ently underreports high levels of indels because of the hetero-
light-inducible heterodimerizing proteins to dCas9 (CIB1-dCas9) duplex dependency of the Surveyor enzyme (Vouillot et al.,
and a transcriptional activator (CRY2-VP64) provides for opto- 2015). Moreover, Surveyor quantification by gel image analysis
genetically controlled transcriptional activation (Polstein and or electrophoresis devices like the Qiagen Qiaxcel or Agilent
Gersbach, 2015). Although background activity remains to be TapeStation has limited sensitivity compared with targeted
optimized, this approach could facilitate effective manipulation deep sequencing. As such, targeted sequencing should become
of the spatial and temporal specificity of a dCas9 synthetic tran- the standard assay for validating off-target events.
scription factor. Targeted deep sequencing of off-target events is particularly
amenable to standardization and reporting of developments in
The Need for Standardization the field, although the method is variably applied across the liter-
As demonstrated, the field of Cas9 specificity is developing ature in terms of experimental execution and computational
rapidly and has already accomplished dramatic improvements analysis. In our view, it is important to minimize PCR cycle num-
over the first-generation methods. In thinking about developing ber, use error-correcting DNA polymerases, consider the
medicines, it is important to keep in mind that all methods for sequencing coverage per input genome copy, and compare
manipulating therapeutic targets, ranging from genome editing reads with appropriate reference sequences that take into ac-
reagents and siRNAs to antibodies and small molecules, are count genetic variation from the standard reference genomes
prone to unintended off-target effects. The expectation for the (Box 1). Many groups currently use custom analysis pipelines,
specificity of biological engineering and Cas9 in particular should so raw data and code should be shared when possible via web
be continuous improvement and appropriate vigilance. Striving repositories such as Github. Ideally, robust analysis tools will

Molecular Cell 63, August 4, 2016 365


Molecular Cell

Review

Box 2. Recommended Practices for Minimizing, Identifying, and Measuring Cas9 Off-Target Cleavage

1. Important: Design guides for uniqueness in the appropriate reference genome


a. No off-by-one or off-by-two mismatch sites
b. Highly ranked by specificity score (Hsu et al., 2013) or CFD (Doench et al., 2016)
2. Preferred: Detect off-target sites with one or more unbiased methods
a. BLESS, GUIDE-seq, Digenome-seq, etc.
3. Important: Validate all detected off-target sites with targeted deep sequencing
a. Prepare library from >10,000 cells
b. Minimize PCR cycle number and use high-fidelity polymerase
c. Determine limit of detection on the basis of ‘‘indel’’ rate in negative control, aim for 0.1% with current technology
4. Iterate steps 2 and 3 to optimize on-target/off-target ratio
a. Titrate dose
b. Use Cas9 variants (eSpCas9, SpCas9-HF)
c. Optimize guide RNA structure (tru-guides, GGX20)

be widely adopted as standard pipelines, as exemplified by the information, in combination with continued progress in off-target
Tuxedo suite (Kim et al., 2013; Trapnell et al., 2013) for RNA- detection methods and Cas9 optimization, could greatly improve
seq. CRISPR-GA (Güell et al., 2014) and CRISPResso (http:// guide selection and enable fair comparison and benchmarking
crispresso.rocks) are two publically available indel analysis across publications while ultimately enabling precise and safe
packages that are available as online or command-line tools. genome editing. Importantly, such standards must be developed
In addition, standard guide RNAs for specificity experiments with a focus on maximizing end-user adoption rates, minimizing
already help researchers compare methods and should continue usability hurdles, and incorporating user feedback, because ‘‘a
to be used as benchmarks. For example, each of the BLESS, standard that is not widely used is not really a standard’’ (Cher-
HTGTS, GUIDE-seq, IDLV capture, and Digenome-seq papers vitz et al., 2011). Historically, working group scientific consor-
assesses at least one shared guide RNA (targeting either tiums have led the development of ‘‘Omics’’ standards while sci-
EMX1 or VEGFA). The different methods identified intersecting entific journals and funding agencies have been critical in
but unequal sets of off-targets for the standard guide RNAs, con- enforcing their broad compliance. Aligning genome editing
firming that none of the unbiased methods are yet perfectly sen- stakeholders toward these goals would represent a discrete
sitive and reliable. However, the off-target sites detected across step forward in the maturation of this field.
multiple studies are also generally the most frequently detected
and edited, whereas the others are very rare, an encouraging Conclusions
result for the field at large (Gabriel et al., 2015). The scientific landscape around Cas9 specificity is rapidly
We propose that a set of guidelines for reporting CRISPR- evolving, with marked improvements recently achieved in guide
Cas9 specificity research should be established, much like the selection, protein and guide engineering, off-target detection
minimum information for publication of quantitative real-time methods, and more. Notably, two recent protein engineering ap-
PCR experiments guidelines for qPCR (Bustin et al., 2009), the proaches (eSpCas9 and SpCas9-HF) have elegantly increased
minimum information about a microarray experiment for micro- SpCas9 specificity, most likely by reducing tolerance for mis-
arrays (Brazma et al., 2001), the minimum information about a matched DNA binding, although their overall on-target efficiency
proteomics experiment for proteomics (Taylor, 2006; Taylor relative to the wild-type enzyme remain to be studied by more
et al., 2007; Martı́nez-Bartolomé et al., 2014), the minimum infor- groups in different applications and organisms. These highly spe-
mation specification for in situ hybridization and immunohisto- cific variants should encourage development of refined off-target
chemistry experiments for fluorescence in situ hybridization ex- detection methods with limits of detection well below 0.1%. Cas9
periments (Deutsch et al., 2006, 2008), as well as others (Chervitz orthologs and non-Cas9 CRISPR enzymes are also emerging as
et al., 2011). Such guidelines are directed toward ensuring un- genome editing tools, but will ultimately face the same challenges
ambiguous communication of experimental designs and results, (Shmakov et al., 2015; Zetsche et al., 2015). In order to propagate
such that independent scientists can fairly interpret their conclu- these advances throughout the broader research community, the
sions in the context of the broader field while reproducing those genome editing field will greatly benefit from discussion and
results without requiring additional information. These ‘‘minimum consensus about best practices for future research (Box 2) as
information’’ requirements should include the number of off- well as data standardization. Beyond specificity, key challenges
target sites assessed, the off-target prediction constraints, cell and opportunities for the field lie in improving delivery and
type, delivery method, time point, dosage, analysis method, increasing efficiency, especially for precise edits.
computational tools, sample size, and other key factors known
to impact specificity.
ACKNOWLEDGMENTS
The research community could begin to report guide RNA on-
and off-target activity to shared databases, such as EENdb (Xiao We gratefully acknowledge Alexandra Glucksmann, Hari Jayaram, and An-
et al., 2013) or CrisprGE (Kaur et al., 2015). Such standardized drew Hack for expert comments on the article. J.T. and V.E.M. are employees

366 Molecular Cell 63, August 4, 2016


Molecular Cell

Review
and P.D.H. is a former employee of Editas Medicine, which is developing ther- Cho, S.W., Kim, S., Kim, J.M., and Kim, J.S. (2013a). Targeted genome engi-
apeutics based on CRISPR-Cas9 genome editing technology. P.D.H. is sup- neering in human cells with the Cas9 RNA-guided endonuclease. Nat. Bio-
ported by the NIH (1DP5OD021369-01) and the Helmsley Charitable Trust. technol. 31, 230–232.

Cho, S.W., Lee, J., Carroll, D., Kim, J.S., and Lee, J. (2013b). Heritable gene
REFERENCES knockout in Caenorhabditis elegans by direct injection of Cas9-sgRNA ribonu-
cleoproteins. Genetics 195, 1177–1180.
Abudayyeh, O.O., Gootenberg, J.S., Konermann, S., Joung, J., Slaymaker, Cho, S.W., Kim, S., Kim, Y., Kweon, J., Kim, H.S., Bae, S., and Kim, J.S. (2014).
I.M., Cox, D.B., Shmakov, S., Makarova, K.S., Semenova, E., Minakhin, L., Analysis of off-target effects of CRISPR/Cas-derived RNA-guided endonucle-
et al. (2016). C2c2 is a single-component programmable RNA-guided RNA- ases and nickases. Genome Res. 24, 132–141.
targeting CRISPR effector. Science, Published online June 2, 2016. http://
dx.doi.org/10.1126/science.aaf5573. Christian, M., Cermak, T., Doyle, E.L., Schmidt, C., Zhang, F., Hummel, A.,
Bogdanove, A.J., and Voytas, D.F. (2010). Targeting DNA double-strand
Alic, N., Ayoub, N., Landrieux, E., Favry, E., Baudouin-Cornu, P., Riva, M., and breaks with TAL effector nucleases. Genetics 186, 757–761.
Carles, C. (2007). Selectivity and proofreading both contribute significantly to
the fidelity of RNA polymerase III transcription. Proc. Natl. Acad. Sci. U S A Chylinski, K., Le Rhun, A., and Charpentier, E. (2013). The tracrRNA and Cas9
104, 10400–10405. families of type II CRISPR-Cas immunity systems. RNA Biol. 10, 726–737.

Anders, C., Niewoehner, O., Duerst, A., and Jinek, M. (2014). Structural basis Cong, L., Ran, F.A., Cox, D., Lin, S., Barretto, R., Habib, N., Hsu, P.D., Wu, X.,
of PAM-dependent target DNA recognition by the Cas9 endonuclease. Nature Jiang, W., Marraffini, L.A., and Zhang, F. (2013). Multiplex genome engineering
513, 569–573. using CRISPR/Cas systems. Science 339, 819–823.

Barrangou, R., Fremaux, C., Deveau, H., Richards, M., Boyaval, P., Moineau, Cox, D.B., Platt, R.J., and Zhang, F. (2015). Therapeutic genome editing: pros-
S., Romero, D.A., and Horvath, P. (2007). CRISPR provides acquired resis- pects and challenges. Nat. Med. 21, 121–131.
tance against viruses in prokaryotes. Science 315, 1709–1712.
Dahlman, J.E., Abudayyeh, O.O., Joung, J., Gootenberg, J.S., Zhang, F., and
Bibikova, M., Carroll, D., Segal, D.J., Trautman, J.K., Smith, J., Kim, Y.G., and Konermann, S. (2015). Orthogonal gene knockout and activation with a cata-
Chandrasegaran, S. (2001). Stimulation of homologous recombination through lytically active Cas9 nuclease. Nat. Biotechnol. 33, 1159–1161.
targeted cleavage by chimeric nucleases. Mol. Cell. Biol. 21, 289–297.
Dang, Y., Jia, G., Choi, J., Ma, H., Anaya, E., Ye, C., Shankar, P., and Wu, H.
Boch, J., Scholze, H., Schornack, S., Landgraf, A., Hahn, S., Kay, S., Lahaye, (2015). Optimizing sgRNA structure to improve CRISPR-Cas9 knockout effi-
T., Nickstadt, A., and Bonas, U. (2009). Breaking the code of DNA binding ciency. Genome Biol. 16, 280.
specificity of TAL-type III effectors. Science 326, 1509–1512.
Davis, K.M., Pattanayak, V., Thompson, D.B., Zuris, J.A., and Liu, D.R. (2015).
Bolukbasi, M.F., Gupta, A., Oikemus, S., Derr, A.G., Garber, M., Brodsky, Small molecule-triggered Cas9 protein with improved genome-editing speci-
M.H., Zhu, L.J., and Wolfe, S.A. (2015). DNA-binding-domain fusions enhance ficity. Nat. Chem. Biol. 11, 316–318.
the targeting range and precision of Cas9. Nat. Methods 12, 1150–1156.
Deltcheva, E., Chylinski, K., Sharma, C.M., Gonzales, K., Chao, Y., Pirzada,
Bolukbasi, M.F., Gupta, A., and Wolfe, S.A. (2016). Creating and evaluating ac- Z.A., Eckert, M.R., Vogel, J., and Charpentier, E. (2011). CRISPR RNA matura-
curate CRISPR-Cas9 scalpels for genomic surgery. Nat. Methods 13, 41–50. tion by trans-encoded small RNA and host factor RNase III. Nature 471,
602–607.
Brazma, A., Hingamp, P., Quackenbush, J., Sherlock, G., Spellman, P.,
Deutsch, E.W., Ball, C.A., Bova, G.S., Brazma, A., Bumgarner, R.E., Campbell,
Stoeckert, C., Aach, J., Ansorge, W., Ball, C.A., Causton, H.C., et al. (2001).
D., Causton, H.C., Christiansen, J., Davidson, D., Eichner, L.J., et al. (2006).
Minimum information about a microarray experiment (MIAME)—toward stan-
Development of the minimum information specification for in situ hybridization
dards for microarray data. Nat. Genet. 29, 365–371.
and immunohistochemistry experiments (MISFISHIE). OMICS 10, 205–208.
Briner, A.E., Donohoue, P.D., Gomaa, A.A., Selle, K., Slorach, E.M., Nye, C.H.,
Deutsch, E.W., Ball, C.A., Berman, J.J., Bova, G.S., Brazma, A., Bumgarner,
Haurwitz, R.E., Beisel, C.L., May, A.P., and Barrangou, R. (2014). Guide RNA
R.E., Campbell, D., Causton, H.C., Christiansen, J.H., Daian, F., et al. (2008).
functional modules direct Cas9 activity and orthogonality. Mol. Cell 56,
Minimum information specification for in situ hybridization and immunohisto-
333–339.
chemistry experiments (MISFISHIE). Nat. Biotechnol. 26, 305–312.
Brouns, S.J., Jore, M.M., Lundgren, M., Westra, E.R., Slijkhuis, R.J., Snijders, Doench, J.G., Hartenian, E., Graham, D.B., Tothova, Z., Hegde, M., Smith, I.,
A.P., Dickman, M.J., Makarova, K.S., Koonin, E.V., and van der Oost, J. (2008). Sullender, M., Ebert, B.L., Xavier, R.J., and Root, D.E. (2014). Rational design
Small CRISPR RNAs guide antiviral defense in prokaryotes. Science 321, of highly active sgRNAs for CRISPR-Cas9-mediated gene inactivation. Nat.
960–964. Biotechnol. 32, 1262–1267.
Bustin, S.A., Benes, V., Garson, J.A., Hellemans, J., Huggett, J., Kubista, M., Doench, J.G., Fusi, N., Sullender, M., Hegde, M., Vaimberg, E.W., Donovan,
Mueller, R., Nolan, T., Pfaffl, M.W., Shipley, G.L., et al. (2009). The MIQE guide- K.F., Smith, I., Tothova, Z., Wilen, C., Orchard, R., et al. (2016). Optimized
lines: minimum information for publication of quantitative real-time PCR exper- sgRNA design to maximize activity and minimize off-target effects of
iments. Clin. Chem. 55, 611–622. CRISPR-Cas9. Nat. Biotechnol. 34, 184–191.
Chari, R., Mali, P., Moosburner, M., and Church, G.M. (2015). Unraveling Fonfara, I., Le Rhun, A., Chylinski, K., Makarova, K.S., Lécrivain, A.L.,
CRISPR-Cas9 genome engineering parameters via a library-on-library Bzdrenga, J., Koonin, E.V., and Charpentier, E. (2014). Phylogeny of Cas9 de-
approach. Nat. Methods 12, 823–826. termines functional exchangeability of dual-RNA and Cas9 among ortholo-
gous type II CRISPR-Cas systems. Nucleic Acids Res. 42, 2577–2590.
Chen, B., Gilbert, L.A., Cimini, B.A., Schnitzbauer, J., Zhang, W., Li, G.W.,
Park, J., Blackburn, E.H., Weissman, J.S., Qi, L.S., and Huang, B. (2013). Dy- , M., Le Rhun, A., and Charpentier, E. (2016).
Fonfara, I., Richter, H., Bratovic
namic imaging of genomic loci in living human cells by an optimized CRISPR/ The CRISPR-associated DNA-cleaving enzyme Cpf1 also processes precur-
Cas system. Cell 155, 1479–1491. sor CRISPR RNA. Nature 532, 517–521.

Chen, B., Hu, J., Almeida, R., Liu, H., Balakrishnan, S., Covill-Cooke, C., Lim, Friedland, A.E., Baral, R., Singhal, P., Loveluck, K., Shen, S., Sanchez, M.,
W.A., and Huang, B. (2016). Expanding the CRISPR imaging toolset with Marco, E., Gotta, G.M., Maeder, M.L., Kennedy, E.M., et al. (2015). Character-
Staphylococcus aureus Cas9 for simultaneous imaging of multiple genomic ization of Staphylococcus aureus Cas9: a smaller Cas9 for all-in-one adeno-
loci. Nucleic Acids Res. 44, e75. associated virus delivery and paired nickase applications. Genome Biol. 16,
257.
Chervitz, S.A., Deutsch, E.W., Field, D., Parkinson, H., Quackenbush, J.,
Rocca-Serra, P., Sansone, S.A., Stoeckert, C.J., Jr., Taylor, C.F., Taylor, R., Frock, R.L., Hu, J., Meyers, R.M., Ho, Y.J., Kii, E., and Alt, F.W. (2015).
and Ball, C.A. (2011). Data standards for Omics data: the basis of data sharing Genome-wide detection of DNA double-stranded breaks induced by engi-
and reuse. Methods Mol. Biol. 719, 31–69. neered nucleases. Nat. Biotechnol. 33, 179–186.

Molecular Cell 63, August 4, 2016 367


Molecular Cell

Review
Fu, Y., Foden, J.A., Khayter, C., Maeder, M.L., Reyon, D., Joung, J.K., and Jiang, F., Taylor, D.W., Chen, J.S., Kornfeld, J.E., Zhou, K., Thompson, A.J.,
Sander, J.D. (2013). High-frequency off-target mutagenesis induced by Nogales, E., and Doudna, J.A. (2016). Structures of a CRISPR-Cas9 R-loop
CRISPR-Cas nucleases in human cells. Nat. Biotechnol. 31, 822–826. complex primed for DNA cleavage. Science 351, 867–871.

Fu, B.X., Hansen, L.L., Artiles, K.L., Nonet, M.L., and Fire, A.Z. (2014a). Land- Jinek, M., Chylinski, K., Fonfara, I., Hauer, M., Doudna, J.A., and Charpentier,
scape of target:guide homology effects on Cas9-mediated cleavage. Nucleic E. (2012). A programmable dual-RNA-guided DNA endonuclease in adaptive
Acids Res. 42, 13778–13787. bacterial immunity. Science 337, 816–821.

Fu, Y., Sander, J.D., Reyon, D., Cascio, V.M., and Joung, J.K. (2014b). Joung, J.K. (2015). Unwanted mutations: standards needed for gene-editing
Improving CRISPR-Cas nuclease specificity using truncated guide RNAs. errors. Nature 523, 158.
Nat. Biotechnol. 32, 279–284.
Kaur, K., Tandon, H., Gupta, A.K., and Kumar, M. (2015). CrisprGE: a central
Gabriel, R., von Kalle, C., and Schmidt, M. (2015). Mapping the precision of hub of CRISPR/Cas-based genome editing. Database (Oxford) 2015, bav055.
genome editing. Nat. Biotechnol. 33, 150–152.
Kiani, S., Chavez, A., Tuttle, M., Hall, R.N., Chari, R., Ter-Ovanesyan, D., Qian,
Gasiunas, G., Barrangou, R., Horvath, P., and Siksnys, V. (2012). Cas9-crRNA
J., Pruitt, B.W., Beal, J., Vora, S., et al. (2015). Cas9 gRNA engineering for
ribonucleoprotein complex mediates specific DNA cleavage for adaptive im-
genome editing, activation and repression. Nat. Methods 12, 1051–1054.
munity in bacteria. Proc. Natl. Acad. Sci. U S A 109, E2579–E2586.
Kim, D., Pertea, G., Trapnell, C., Pimentel, H., Kelley, R., and Salzberg, S.L.
Ge, Y., Powell, S., Van Roey, M., and McArthur, J.G. (2001). Factors influ-
(2013). TopHat2: accurate alignment of transcriptomes in the presence of in-
encing the development of an anti-factor IX (FIX) immune response following
sertions, deletions and gene fusions. Genome Biol. 14, R36.
administration of adeno-associated virus-FIX. Blood 97, 3733–3737.

Gilbert, L.A., Horlbeck, M.A., Adamson, B., Villalta, J.E., Chen, Y., Whitehead, Kim, S., Kim, D., Cho, S.W., Kim, J., and Kim, J.S. (2014). Highly efficient RNA-
E.H., Guimaraes, C., Panning, B., Ploegh, H.L., Bassik, M.C., et al. (2014). guided genome editing in human cells via delivery of purified Cas9 ribonucleo-
Genome-scale CRISPR-mediated control of gene repression and activation. proteins. Genome Res. 24, 1012–1019.
Cell 159, 647–661.
Kim, D., Bae, S., Park, J., Kim, E., Kim, S., Yu, H.R., Hwang, J., Kim, J.I., and
Gori, J.L., Hsu, P.D., Maeder, M.L., Shen, S., Welstead, G.G., and Bumcrot, D. Kim, J.S. (2015). Digenome-seq: genome-wide profiling of CRISPR-Cas9 off-
(2015). Delivery and specificity of CRISPR-Cas9 genome editing technologies target effects in human cells. Nat. Methods 12, 237–243, 1, 243.
for human gene therapy. Hum. Gene Ther. 26, 443–451.
Kim, D., Kim, J., Hur, J.K., Been, K.W., Yoon, S.H., and Kim, J.S. (2016a).
Güell, M., Yang, L., and Church, G.M. (2014). Genome editing assessment us- Genome-wide analysis reveals specificities of Cpf1 endonucleases in human
ing CRISPR Genome Analyzer (CRISPR-GA). Bioinformatics 30, 2968–2970. cells. Nat. Biotechnol., Published online June 6, 2016. http://dx.doi.org/10.
1038/nbt.3609.
Guilinger, J.P., Thompson, D.B., and Liu, D.R. (2014). Fusion of catalytically
inactive Cas9 to FokI nuclease improves the specificity of genome modifica- Kim, D., Kim, S., Kim, S., Park, J., and Kim, J.S. (2016b). Genome-wide target
tion. Nat. Biotechnol. 32, 577–582. specificities of CRISPR-Cas9 nucleases revealed by multiplex Digenome-seq.
Genome Res. 26, 406–415.
Guschin, D.Y., Waite, A.J., Katibah, G.E., Miller, J.C., Holmes, M.C., and
Rebar, E.J. (2010). A rapid and general assay for monitoring endogenous Kleinstiver, B.P., Prew, M.S., Tsai, S.Q., Nguyen, N.T., Topkar, V.V., Zheng, Z.,
gene modification. Methods Mol. Biol. 649, 247–256. and Joung, J.K. (2015a). Broadening the targeting range of Staphylococcus
aureus CRISPR-Cas9 by modifying PAM recognition. Nat. Biotechnol. 33,
Hastie, E., and Samulski, R.J. (2015). Adeno-associated virus at 50: a golden 1293–1298.
anniversary of discovery, research, and gene therapy success—a personal
perspective. Hum. Gene Ther. 26, 257–265. Kleinstiver, B.P., Prew, M.S., Tsai, S.Q., Topkar, V.V., Nguyen, N.T., Zheng, Z.,
Gonzales, A.P., Li, Z., Peterson, R.T., Yeh, J.R., et al. (2015b). Engineered
Hendel, A., Bak, R.O., Clark, J.T., Kennedy, A.B., Ryan, D.E., Roy, S., Stein- CRISPR-Cas9 nucleases with altered PAM specificities. Nature 523, 481–485.
feld, I., Lunstad, B.D., Kaiser, R.J., Wilkens, A.B., et al. (2015). Chemically
modified guide RNAs enhance CRISPR-Cas genome editing in human primary Kleinstiver, B.P., Pattanayak, V., Prew, M.S., Tsai, S.Q., Nguyen, N.T., Zheng,
cells. Nat. Biotechnol. 33, 985–989. Z., and Joung, J.K. (2016a). High-fidelity CRISPR-Cas9 nucleases with no
detectable genome-wide off-target effects. Nature 529, 490–495.
Hou, Z., Zhang, Y., Propson, N.E., Howden, S.E., Chu, L.F., Sontheimer, E.J.,
and Thomson, J.A. (2013). Efficient genome engineering in human pluripotent Kleinstiver, B.P., Tsai, S.Q., Prew, M.S., Nguyen, N.T., Welch, M.M., Lopez,
stem cells using Cas9 from Neisseria meningitidis. Proc. Natl. Acad. Sci. U S A J.M., McCaw, Z.R., Aryee, M.J., and Joung, J.K. (2016b). Genome-wide spec-
110, 15644–15649. ificities of CRISPR-Cas Cpf1 nucleases in human cells. Nat. Biotechnol., Pub-
lished online June 27, 2016. http://dx.doi.org/10.1038/nbt.3620.
Hsu, P.D., Scott, D.A., Weinstein, J.A., Ran, F.A., Konermann, S., Agarwala, V.,
Li, Y., Fine, E.J., Wu, X., Shalem, O., et al. (2013). DNA targeting specificity of
Knight, S.C., Xie, L., Deng, W., Guglielmi, B., Witkowsky, L.B., Bosanac, L.,
RNA-guided Cas9 nucleases. Nat. Biotechnol. 31, 827–832.
Zhang, E.T., El Beheiry, M., Masson, J.B., Dahan, M., et al. (2015). Dynamics
of CRISPR-Cas9 genome interrogation in living cells. Science 350, 823–826.
Hsu, P.D., Lander, E.S., and Zhang, F. (2014). Development and applications
of CRISPR-Cas9 for genome engineering. Cell 157, 1262–1278.
Konermann, S., Brigham, M.D., Trevino, A.E., Hsu, P.D., Heidenreich, M.,
Huang, J., Brieba, L.G., and Sousa, R. (2000). Misincorporation by wild-type Cong, L., Platt, R.J., Scott, D.A., Church, G.M., and Zhang, F. (2013). Optical
and mutant T7 RNA polymerases: identification of interactions that reduce control of mammalian endogenous transcription and epigenetic states. Nature
misincorporation rates by stabilizing the catalytically incompetent open 500, 472–476.
conformation. Biochemistry 39, 11571–11580.
Konermann, S., Brigham, M.D., Trevino, A.E., Joung, J., Abudayyeh, O.O.,
Jansen, R., Embden, J.D., Gaastra, W., and Schouls, L.M. (2002). Identifica- Barcena, C., Hsu, P.D., Habib, N., Gootenberg, J.S., Nishimasu, H., et al.
tion of genes that are associated with DNA repeats in prokaryotes. Mol. Micro- (2015). Genome-scale transcriptional activation by an engineered CRISPR-
biol. 43, 1565–1575. Cas9 complex. Nature 517, 583–588.

Jiang, W., Bikard, D., Cox, D., Zhang, F., and Marraffini, L.A. (2013). RNA- Koo, T., Lee, J., and Kim, J.S. (2015). Measuring and reducing off-target activ-
guided editing of bacterial genomes using CRISPR-Cas systems. Nat. Bio- ities of programmable nucleases including CRISPR-Cas9. Mol. Cells 38,
technol. 31, 233–239. 475–481.

Jiang, F., Zhou, K., Ma, L., Gressel, S., and Doudna, J.A. (2015). Structural Kuscu, C., Arslan, S., Singh, R., Thorpe, J., and Adli, M. (2014). Genome-wide
biology. A Cas9-guide RNA complex preorganized for target DNA recognition. analysis reveals characteristics of off-target sites bound by the Cas9 endonu-
Science 348, 1477–1481. clease. Nat. Biotechnol. 32, 677–683.

368 Molecular Cell 63, August 4, 2016


Molecular Cell

Review
Lee, C.M., Cradick, T.J., and Bao, G. (2016). The Neisseria meningitidis Nishimasu, H., Cong, L., Yan, W.X., Ran, F.A., Zetsche, B., Li, Y., Kurabayashi,
CRISPR-Cas9 system enables specific genome editing in mammalian cells. A., Ishitani, R., Zhang, F., and Nureki, O. (2015). Crystal structure of Staphylo-
Mol. Ther. 24, 645–654. coccus aureus Cas9. Cell 162, 1113–1126.

Limberis, M.P., and Wilson, J.M. (2006). Adeno-associated virus serotype 9 Nissim, L., Perli, S.D., Fridkin, A., Perez-Pinera, P., and Lu, T.K. (2014). Multi-
vectors transduce murine alveolar and nasal epithelia and can be readminis- plexed and programmable regulation of gene networks with an integrated RNA
tered. Proc. Natl. Acad. Sci. U S A 103, 12993–12998. and CRISPR/Cas toolkit in human cells. Mol. Cell 54, 698–710.

Lin, S., Staahl, B.T., Alla, R.K., and Doudna, J.A. (2014). Enhanced homology- Pattanayak, V., Lin, S., Guilinger, J.P., Ma, E., Doudna, J.A., and Liu, D.R.
directed human genome engineering by controlled timing of CRISPR/Cas9 de- (2013). High-throughput profiling of off-target DNA cleavage reveals RNA-pro-
livery. eLife 3, e04766. grammed Cas9 nuclease specificity. Nat. Biotechnol. 31, 839–843.

Ma, E., Harrington, L.B., O’Connell, M.R., Zhou, K., and Doudna, J.A. (2015). Perez-Pinera, P., Kocak, D.D., Vockley, C.M., Adler, A.F., Kabadi, A.M., Pol-
Single-stranded DNA cleavage by divergent CRISPR-Cas9 enzymes. Mol. stein, L.R., Thakore, P.I., Glass, K.A., Ousterout, D.G., Leong, K.W., et al.
Cell 60, 398–407. (2013). RNA-guided gene activation by CRISPR-Cas9-based transcription
factors. Nat. Methods 10, 973–976.
MacPherson, C. (2015). The CRISPR Software Matchmaker: a new tool for
Polstein, L.R., and Gersbach, C.A. (2015). A light-inducible CRISPR-Cas9 sys-
choosing the best CRISPR software for your needs. Addgene, http://blog.
tem for control of endogenous gene activation. Nat. Chem. Biol. 11, 198–200.
addgene.org/the-crispr-software-matchmaker-a-new-tool-for-choosing-the-
best-crispr-software-for-your-needs.
Polstein, L.R., Perez-Pinera, P., Kocak, D.D., Vockley, C.M., Bledsoe, P.,
Song, L., Safi, A., Crawford, G.E., Reddy, T.E., and Gersbach, C.A. (2015).
Majowicz, A., Maczuga, P., Kwikkers, K.L., van der Marel, S., van Logtenstein,
Genome-wide specificity of DNA binding, gene regulation, and chromatin re-
R., Petry, H., van Deventer, S.J., Konstantinova, P., and Ferreira, V. (2013). Mir-
modeling by TALE- and CRISPR/Cas9-based transcriptional activators.
142-3p target sequences reduce transgene-directed immunogenicity
Genome Res. 25, 1158–1169.
following intramuscular adeno-associated virus 1 vector-mediated gene deliv-
ery. J. Gene Med. 15, 219–232. Porteus, M. (2016). Genome editing: a new approach to human therapeutics.
Annu. Rev. Pharmacol. Toxicol. 56, 163–190.
Makarova, K.S., and Koonin, E.V. (2015). Annotation and classification of
CRISPR-Cas systems. Methods Mol. Biol. 1311, 47–75. Pourcel, C., Salvignol, G., and Vergnaud, G. (2005). CRISPR elements in Yer-
sinia pestis acquire new repeats by preferential uptake of bacteriophage DNA,
Mali, P., Aach, J., Stranges, P.B., Esvelt, K.M., Moosburner, M., Kosuri, S., and provide additional tools for evolutionary studies. Microbiology 151,
Yang, L., and Church, G.M. (2013a). CAS9 transcriptional activators for target 653–663.
specificity screening and paired nickases for cooperative genome engineer-
ing. Nat. Biotechnol. 31, 833–838. Rahdar, M., McMahon, M.A., Prakash, T.P., Swayze, E.E., Bennett, C.F., and
Cleveland, D.W. (2015). Synthetic CRISPR RNA-Cas9-guided genome editing
Mali, P., Esvelt, K.M., and Church, G.M. (2013b). Cas9 as a versatile tool for in human cells. Proc. Natl. Acad. Sci. U S A 112, E7110–E7117.
engineering biology. Nat. Methods 10, 957–963.
Ran, F.A., Hsu, P.D., Lin, C.Y., Gootenberg, J.S., Konermann, S., Trevino, A.E.,
Mali, P., Yang, L., Esvelt, K.M., Aach, J., Guell, M., DiCarlo, J.E., Norville, J.E., Scott, D.A., Inoue, A., Matoba, S., Zhang, Y., and Zhang, F. (2013a). Double
and Church, G.M. (2013c). RNA-guided human genome engineering via Cas9. nicking by RNA-guided CRISPR Cas9 for enhanced genome editing speci-
Science 339, 823–826. ficity. Cell 154, 1380–1389.

Martı́nez-Bartolomé, S., Binz, P.A., and Albar, J.P. (2014). The minimal infor- Ran, F.A., Hsu, P.D., Wright, J., Agarwala, V., Scott, D.A., and Zhang, F.
mation about a proteomics experiment (MIAPE) from the Proteomics Stan- (2013b). Genome engineering using the CRISPR-Cas9 system. Nat. Protoc.
dards Initiative. Methods Mol. Biol. 1072, 765–780. 8, 2281–2308.

Maruyama, T., Dougan, S.K., Truttmann, M.C., Bilate, A.M., Ingram, J.R., and Ran, F.A., Cong, L., Yan, W.X., Scott, D.A., Gootenberg, J.S., Kriz, A.J., Zet-
Ploegh, H.L. (2015). Increasing the efficiency of precise genome editing with sche, B., Shalem, O., Wu, X., Makarova, K.S., et al. (2015). In vivo genome ed-
CRISPR-Cas9 by inhibition of nonhomologous end joining. Nat. Biotechnol. iting using Staphylococcus aureus Cas9. Nature 520, 186–191.
33, 538–542.
Ranganathan, V., Wahlin, K., Maruotti, J., and Zack, D.J. (2014). Expansion of
Miller, J.C., Holmes, M.C., Wang, J., Guschin, D.Y., Lee, Y.L., Rupniewski, I., the CRISPR-Cas9 genome targeting space through the use of H1 promoter-
Beausejour, C.M., Waite, A.J., Wang, N.S., Kim, K.A., et al. (2007). An expressed guide RNAs. Nat. Commun. 5, 4516.
improved zinc-finger nuclease architecture for highly specific genome editing.
Nat. Biotechnol. 25, 778–785. Richardson, C.D., Ray, G.J., DeWitt, M.A., Curie, G.L., and Corn, J.E. (2016).
Enhancing homology-directed genome editing by catalytically active and inac-
tive CRISPR-Cas9 using asymmetric donor DNA. Nat. Biotechnol. 34,
Miller, J.C., Tan, S., Qiao, G., Barlow, K.A., Wang, J., Xia, D.F., Meng, X., Pa-
339–344.
schon, D.E., Leung, E., Hinkley, S.J., et al. (2011). A TALE nuclease architec-
ture for efficient genome editing. Nat. Biotechnol. 29, 143–148.
Sander, J.D., and Joung, J.K. (2014). CRISPR-Cas systems for editing, regu-
lating and targeting genomes. Nat. Biotechnol. 32, 347–355.
Mojica, F.J., Dı́ez-Villaseñor, C., Soria, E., and Juez, G. (2000). Biological sig-
nificance of a family of regularly spaced repeats in the genomes of Archaea, Sapranauskas, R., Gasiunas, G., Fremaux, C., Barrangou, R., Horvath, P., and
Bacteria and mitochondria. Mol. Microbiol. 36, 244–246. Siksnys, V. (2011). The Streptococcus thermophilus CRISPR/Cas system pro-
vides immunity in Escherichia coli. Nucleic Acids Res. 39, 9275–9282.
Mojica, F.J., Dı́ez-Villaseñor, C., Garcı́a-Martı́nez, J., and Soria, E. (2005).
Intervening sequences of regularly spaced prokaryotic repeats derive from Shalem, O., Sanjana, N.E., Hartenian, E., Shi, X., Scott, D.A., Mikkelsen, T.S.,
foreign genetic elements. J. Mol. Evol. 60, 174–182. Heckl, D., Ebert, B.L., Root, D.E., Doench, J.G., and Zhang, F. (2014).
Genome-scale CRISPR-Cas9 knockout screening in human cells. Science
Moscou, M.J., and Bogdanove, A.J. (2009). A simple cipher governs DNA 343, 84–87.
recognition by TAL effectors. Science 326, 1501.
Shen, B., Zhang, W., Zhang, J., Zhou, J., Wang, J., Chen, L., Wang, L., Hodg-
Müller, M., Lee, C.M., Gasiunas, G., Davis, T.H., Cradick, T.J., Siksnys, V., kins, A., Iyer, V., Huang, X., and Skarnes, W.C. (2014). Efficient genome modi-
Bao, G., Cathomen, T., and Mussolino, C. (2016). Streptococcus thermophilus fication by CRISPR-Cas9 nickase with minimal off-target effects. Nat.
CRISPR-Cas9 systems enable specific editing of the human genome. Mol. Methods 11, 399–402.
Ther. 24, 636–644.
Shmakov, S., Abudayyeh, O.O., Makarova, K.S., Wolf, Y.I., Gootenberg, J.S.,
Nishimasu, H., Ran, F.A., Hsu, P.D., Konermann, S., Shehata, S.I., Dohmae, Semenova, E., Minakhin, L., Joung, J., Konermann, S., Severinov, K., et al.
N., Ishitani, R., Zhang, F., and Nureki, O. (2014). Crystal structure of Cas9 in (2015). Discovery and functional characterization of diverse class 2 CRISPR-
complex with guide RNA and target DNA. Cell 156, 935–949. Cas systems. Mol. Cell 60, 385–397.

Molecular Cell 63, August 4, 2016 369


Molecular Cell

Review
Singh, R., Kuscu, C., Quinlan, A., Qi, Y., and Adli, M. (2015). Cas9-chromatin CRISPR-Cas9 and TALENs using integrase-defective lentiviral vectors. Nat.
binding information enables more accurate CRISPR off-target prediction. Nu- Biotechnol. 33, 175–178.
cleic Acids Res. 43, e118.
Wright, A.V., Nuñez, J.K., and Doudna, J.A. (2016). Biology and applications of
Slaymaker, I.M., Gao, L., Zetsche, B., Scott, D.A., Yan, W.X., and Zhang, F. CRISPR systems: harnessing nature’s toolbox for genome engineering. Cell
(2016). Rationally engineered Cas9 nucleases with improved specificity. Sci- 164, 29–44.
ence 351, 84–88.
Wu, X., Scott, D.A., Kriz, A.J., Chiu, A.C., Hsu, P.D., Dadon, D.B., Cheng, A.W.,
Sternberg, S.H., Redding, S., Jinek, M., Greene, E.C., and Doudna, J.A. (2014). Trevino, A.E., Konermann, S., Chen, S., et al. (2014). Genome-wide binding of
DNA interrogation by the CRISPR RNA-guided endonuclease Cas9. Nature the CRISPR endonuclease Cas9 in mammalian cells. Nat. Biotechnol. 32,
507, 62–67. 670–676.
Sternberg, S.H., LaFrance, B., Kaplan, M., and Doudna, J.A. (2015). Confor-
Wyvekens, N., Topkar, V.V., Khayter, C., Joung, J.K., and Tsai, S.Q. (2015).
mational control of DNA target cleavage by CRISPR-Cas9. Nature 527,
Dimeric CRISPR RNA-guided FokI-dCas9 nucleases directed by truncated
110–113.
gRNAs for highly specific genome editing. Hum. Gene Ther. 26, 425–431.
Sung, Y.H., Kim, J.M., Kim, H.T., Lee, J., Jeon, J., Jin, Y., Choi, J.H., Ban, Y.H.,
Ha, S.J., Kim, C.H., et al. (2014). Highly efficient gene knockout in mice and ze- Xiao, A., Wu, Y., Yang, Z., Hu, Y., Wang, W., Zhang, Y., Kong, L., Gao, G., Zhu,
brafish with RNA-guided endonucleases. Genome Res. 24, 125–131. Z., Lin, S., and Zhang, B. (2013). EENdb: a database and knowledge base of
ZFNs and TALENs for endonuclease engineering. Nucleic Acids Res. 41,
Tabebordbar, M., Zhu, K., Cheng, J.K., Chew, W.L., Widrick, J.J., Yan, W.X., D415–D422.
Maesner, C., Wu, E.Y., Xiao, R., Ran, F.A., et al. (2016). In vivo gene editing
in dystrophic mouse muscle and muscle stem cells. Science 351, 407–411. Xu, H., Xiao, T., Chen, C.H., Li, W., Meyer, C.A., Wu, Q., Wu, D., Cong, L.,
Zhang, F., Liu, J.S., et al. (2015). Sequence determinants of improved CRISPR
Taylor, C.F. (2006). Minimum reporting requirements for proteomics: a MIAPE sgRNA design. Genome Res. 25, 1147–1157.
primer. Proteomics 6 (Suppl 2 ), 39–44.
Yamano, T., Nishimasu, H., Zetsche, B., Hirano, H., Slaymaker, I.M., Li, Y., Fe-
Taylor, C.F., Paton, N.W., Lilley, K.S., Binz, P.A., Julian, R.K., Jr., Jones, A.R., dorova, I., Nakane, T., Makarova, K.S., Koonin, E.V., et al. (2016). Crystal struc-
Zhu, W., Apweiler, R., Aebersold, R., Deutsch, E.W., et al. (2007). The minimum ture of Cpf1 in complex with guide RNA and target DNA. Cell 165, 949–962.
information about a proteomics experiment (MIAPE). Nat. Biotechnol. 25,
887–893. Yang, L., Grishin, D., Wang, G., Aach, J., Zhang, C.Z., Chari, R., Homsy, J., Cai,
X., Zhao, Y., Fan, J.B., et al. (2014). Targeted and genome-wide sequencing
Trapnell, C., Hendrickson, D.G., Sauvageau, M., Goff, L., Rinn, J.L., and reveal single nucleotide variations impacting specificity of Cas9 in human
Pachter, L. (2013). Differential analysis of gene regulation at transcript resolu- stem cells. Nat. Commun. 5, 5507.
tion with RNA-seq. Nat. Biotechnol. 31, 46–53.
Yu, C., Liu, Y., Ma, T., Liu, K., Xu, S., Zhang, Y., Liu, H., La Russa, M., Xie, M.,
Tsai, S.Q., Wyvekens, N., Khayter, C., Foden, J.A., Thapar, V., Reyon, D., Ding, S., and Qi, L.S. (2015). Small molecules enhance CRISPR genome edit-
Goodwin, M.J., Aryee, M.J., and Joung, J.K. (2014). Dimeric CRISPR RNA- ing in pluripotent stem cells. Cell Stem Cell 16, 142–147.
guided FokI nucleases for highly specific genome editing. Nat. Biotechnol.
32, 569–576. Zetsche, B., Gootenberg, J.S., Abudayyeh, O.O., Slaymaker, I.M., Makarova,
K.S., Essletzbichler, P., Volz, S.E., Joung, J., van der Oost, J., Regev, A., et al.
Tsai, S.Q., Zheng, Z., Nguyen, N.T., Liebers, M., Topkar, V.V., Thapar, V., Wy-
(2015). Cpf1 is a single RNA-guided endonuclease of a class 2 CRISPR-Cas
vekens, N., Khayter, C., Iafrate, A.J., Le, L.P., et al. (2015). GUIDE-seq enables
system. Cell 163, 759–771.
genome-wide profiling of off-target cleavage by CRISPR-Cas nucleases. Nat.
Biotechnol. 33, 187–197.
Zhu, L.J., Holmes, B.R., Aronin, N., and Brodsky, M.H. (2014). CRISPRseek: a
Urnov, F.D., Miller, J.C., Lee, Y.L., Beausejour, C.M., Rock, J.M., Augustus, S., bioconductor package to identify target-specific guide RNAs for CRISPR-
Jamieson, A.C., Porteus, M.H., Gregory, P.D., and Holmes, M.C. (2005). Highly Cas9 genome-editing systems. PLoS ONE 9, e108424.
efficient endogenous human gene correction using designed zinc-finger nu-
cleases. Nature 435, 646–651. Zincarelli, C., Soltys, S., Rengo, G., and Rabinowitz, J.E. (2008). Analysis of
AAV serotypes 1-9 mediated gene expression and tropism in mice after sys-
Vouillot, L., Thélie, A., and Pollet, N. (2015). Comparison of T7E1 and surveyor temic injection. Mol. Ther. 16, 1073–1080.
mismatch cleavage assays to detect mutations triggered by engineered nucle-
ases. G3 (Bethesda) 5, 407–415. Zuris, J.A., Thompson, D.B., Shu, Y., Guilinger, J.P., Bessen, J.L., Hu, J.H.,
Maeder, M.L., Joung, J.K., Chen, Z.Y., and Liu, D.R. (2015). Cationic lipid-
Wang, X., Wang, Y., Wu, X., Wang, J., Wang, Y., Qiu, Z., Chang, T., Huang, H., mediated delivery of proteins enables efficient protein-based genome editing
Lin, R.J., and Yee, J.K. (2015). Unbiased detection of off-target cleavage by in vitro and in vivo. Nat. Biotechnol. 33, 73–80.

370 Molecular Cell 63, August 4, 2016

You might also like