Functional interrogation of non-coding DNA through

4 downloads 0 Views 781KB Size Report
Mar 10, 2017 - Bhattacharyya, J.M. Shelton, R. Bassel-Duby, E.N. Olson, Postnatal genome · editing partially restores ..... 2641 (2016) 617–632. M.C. Canver et al. .... the dark road from association to function, Am. J. Hum. Genet. 93 (2013).
Methods 121–122 (2017) 118–129

Contents lists available at ScienceDirect

Methods journal homepage: www.elsevier.com/locate/ymeth

Functional interrogation of non-coding DNA through CRISPR genome editing Matthew C. Canver a, Daniel E. Bauer a,b,c,⇑, Stuart H. Orkin a,b,c,d,⇑ a

Harvard Medical School, Boston, MA 02115, United States Division of Hematology/Oncology, Boston Children’s Hospital, Boston, MA 02115, United States c Department of Pediatric Oncology, Dana-Farber Cancer Institute, Boston, MA 02115, United States d Howard Hughes Medical Institute, Boston, MA 02115, United States b

a r t i c l e

i n f o

Article history: Received 4 January 2017 Received in revised form 18 February 2017 Accepted 3 March 2017 Available online 10 March 2017

a b s t r a c t Methodologies to interrogate non-coding regions have lagged behind coding regions despite comprising the vast majority of the genome. However, the rapid evolution of clustered regularly interspaced short palindromic repeats (CRISPR)-based genome editing has provided a multitude of novel techniques for laboratory investigation including significant contributions to the toolbox for studying non-coding DNA. CRISPR-mediated loss-of-function strategies rely on direct disruption of the underlying sequence or repression of transcription without modifying the targeted DNA sequence. CRISPR-mediated gainof-function approaches similarly benefit from methods to alter the targeted sequence through integration of customized sequence into the genome as well as methods to activate transcription. Here we review CRISPR-based loss- and gain-of-function techniques for the interrogation of non-coding DNA. Ó 2017 Elsevier Inc. All rights reserved.

Contents 1. 2.

3.

4. 5.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Loss of function studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1. Loss of function approaches with DNA sequence modification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1.1. Individual sgRNA targeting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1.2. Dual sgRNA targeting for targeted genomic deletion, inversion, and translocations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1.3. Tiling sgRNAs/Saturating mutagenesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2. Loss of function approaches without DNA sequence modification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Gain of function strategies. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1. Gain of function approaches with DNA sequence modification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2. Gain of function approaches without DNA sequence modification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . General considerations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Author contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Conflict of interest disclosures. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

⇑ Corresponding authors at: Boston Children’s Hospital, 300 Longwood Avenue, Boston, MA 02115, United States. E-mail addresses: [email protected] (D.E. Bauer), [email protected] (S.H. Orkin). http://dx.doi.org/10.1016/j.ymeth.2017.03.008 1046-2023/Ó 2017 Elsevier Inc. All rights reserved.

119 119 120 120 121 121 122 122 123 123 124 124 124 124 124 124

119

M.C. Canver et al. / Methods 121–122 (2017) 118–129

1. Introduction The human genome consists of approximately 3.3 billion base pairs with only 1% protein coding sequence [1–3]. The remaining 99% of the non-coding genome potentially harbors function. However, unlike the coding genome, it lacks mRNA and protein production as functional intermediaries. Decades of research has sought to determine the rules to predict and identify functional elements within the non-coding genome including evolutionary, biochemical/epigenetic, and genetic approaches [2,4]. The understanding of the non-coding genome has been further hampered by a paucity of methods to definitively demonstrate function with reliance on such methods as correlative epigenetic marks [5], heterologous reporter assays including massively parallel reporter assays [6–8], and evolutionary conservation [2,9]. The development of genome editing technologies including zinc finger nucleases (ZFNs) and transcription activator-like effector nucleases (TALENs) offered new methods to identify functional non-coding sequences through loss- and gain-of-function genetic studies [10,11]. The rapid emergence of the clustered regularly interspaced short palindromic repeats (CRISPR)/Cas9 nuclease system as a facile genome editing platform has revolutionized genome editing approaches to study the non-coding genome [12–19]. The CRISPR/Cas9 nuclease system was originally discovered as a prokaryotic adaptive defense system that has been repurposed for eukaryotic genome editing [20]. The CRISPR/Cas9 platform requires an RNA molecule to guide the Cas9 nuclease for sitespecific cleavage upstream of a genomic protospacer adjacent motif (PAM) sequence [12,13]. The endogenous tracrRNA and crRNA have been synthetically fused to create a single chimeric guide RNA (sgRNA) to facilitate genome editing experiments. CRISPR-based editing originally relied on S. pyogenes Cas9 (SpCas9) with an NGG-restricted PAM sequence. However, the CRISPR tool-

box has been expanded to include a wide-range of PAM sequences with many validated for eukaryotic genome editing including engineered S. pyogenes and S. aureus Cas9’s to have altered PAM specificities [12,13,21–27]. Notably, Cpf1 and C2c1 offer alternative CRISPR nucleases to Cas9 as well as uniquely offer site-specific cleavage downstream of PAM sequences [28–30]. CRISPR-based genome editing to interrogate non-coding sequences relies on the endogenous DNA repair pathways that are engaged following double strand break (DSB) induction by a CRISPR-associated nuclease. These repair pathways principally include non-homologous end joining (NHEJ) and homologydirected repair (HDR). The HDR pathway can be exploited by supplying an extrachromosomal repair template to allow for the insertion of customized sequence as opposed to using the sister chromatid as the template for repair [31]. CRISPR-based genome editing systems have offered a host of strategies to study non-coding genomes through both loss-offunction and gain-of-function approaches. In particular, enhancers have been a large focus of study via genome editing [32]. Here we review CRISPR genome editing methodologies for the interrogation of the non-coding genome (Table 1).

2. Loss of function studies CRISPR based loss-of-function studies focus on two different strategies that are distinguished by the requirement to disrupt the underlying sequences. The first strategy employs insertion/ deletion (indel) induction following genomic cleavage to disrupt putative regulatory sequences. Loss-of-function with indel production relies on indel-prone NHEJ repair with a typical deletion spectrum of 1–10 base pairs [12,13,33–35]. The second strategy relies on interference or repression of the underlying sequence without

Table 1 CRISPR approaches to study non-coding DNA. LOF = loss-of-function, GOF = gain-of-function. Approach

# of sgRNA

Strategy

Throughput

Efficiency

Advantages

Disadvantages

sgRNA mutagenesis via indel-prone NHEJ

1

Low-to-high

High

sgRNA with homologous extrachromosomal template for HDR sgRNA with homologous extrachromosomal template for MMEJ sgRNA with nonhomologous plasmid (NHEJ-integration) Targeted deletion, inversion using dual sgRNA

1

Low

Low

Efficient mutagenesis of individual loci; pooled screening Genome customization (e.g., variants, protein tagging)

Narrow regions; often depend on some prior knowledge to design informative targets Not suitable for non-dividing cells; low efficiency; low throughput of loci

Low

Low

Genome customization (e.g., variants, protein tagging)

Not suitable for non-dividing cells; low efficiency; low throughput of loci

Low

High

Low throughput of loci

Low-to-high

Low

Genome customization (eg, variants, protein tagging); high efficiency Large regions; pooled screening

Translocation induction using dual sgRNA

2

Low

Low

Creation of complex genomic rearrangements

Low efficiency; low throughput

Tiling sgRNA/Saturating mutagenesis

>1

High

High

dCas9-KRAB

1

Low-to-high

High

1

Low-to-high

High

dCas9-VP64

1

Low-to-high

High

High-resolution, highthroughput; compatible with any nuclease Reduction without knockout (cell lethal study); pooled screening Reduction without knockout (cell lethal study); pooled screening Pooled screening

Variable resolution as function of underlying sequence PAM availability and off-target potential Dosage; false negatives

dCas9-LSD1

dCas9-p300

1

LOF with genomic modification GOF with genomic modification GOF with genomic modification GOF with genomic modification LOF with genomic modification GOF with genomic modification LOF/GOF with genomic modification LOF without genomic modification LOF without genomic modification GOF without genomic modification GOF without genomic modification

Low-to-high

High

Pooled screening

Dosage; false negatives

1

1

2

Low-resolution; multifocal indels as competing genomic outcome

Dosage; false negatives

Dosage; false negatives

120

M.C. Canver et al. / Methods 121–122 (2017) 118–129

a

b

Individual sgRNA Mutagenesis

5’ 3’

Tiling sgRNA/Saturating Mutagenesis

3’

5’ 3’

3’ 5’

c

Targeted Deletion/Inversion

5’ 5’ 3’

GATA

5’ 3’

3’ 5’

5’ 3’

3’ 5’

3’ 5’

e

d Transcriptional Repression

Individual sgRNA with Homologous Repair Template

KRAB 5’ 3’

GATA GAT ATA

5’ 3’

3’ 5’

5’ 3’

3’ 5’

3’ 5’

GFP

GTTA LSD1 5’ 3’

f

g

Base Editing

GTTA

5’ 3’

3’ 5’

5’ 3’

3’ 5’

h

Homology-Independent NHEJ-Integration

GFP

3’ 5’

Transcriptional Activation

AID

VP64

5’ 3’

C G

3’ 5’

5’

T A

3’

5’ 3’

3’ 5’

p300

3’

5’

5’ 3’

3’ 5’

5’ 3’

3’ 5’

Fig. 1. CRISPR strategies to study non-coding DNA. (a) Individual sgRNA mutagenesis targeting a chromatin mark (shown in red) and a transcription factor binding motif such as GATA. (b) Targeted deletion using a dual sgRNA approach. Deleted sequence is indicated by red dotted lines. Inverted sequence is indicated by red arrows. (c) Tiling sgRNA/ saturating mutagenesis utilizing all sgRNA within genomic loci in a pooled screening format. (d) CRISPR-mediated transcriptional repression. dCas9-KRAB and dCas9-LSD1 are shown. However, other repressive effectors are possible. Chromatin mark is shown in red and SpCas9 is shown in gray. (e) Individual sgRNA with a homologous repair template for such applications as SNP insertion or reporter knock-in. Custom sequences from homologous repair templates are inserted by exploiting HDR or MMEJ. Red sequences indicate homologous sequence and SpCas9 is shown in gray. GFP reporter sequence is shown in green. (f) Base editing. dCas9-AID is shown. However, other base editors are possible. SpCas9 is shown in gray. (g) Homology-independent NHEJ-integration. Shared sequences on plasmid (black circle) and genomic DNA are shown in red. (h) CRISPR-mediated transcriptional activation. dCas9-VP64 and dCas9-p300 are shown. However, other activation effectors are possible. Chromatin mark is shown in red and SpCas9 is shown in gray. (a–h) All schema are shown with SpCas9. The blue lines indicate the 20-mer sgRNA sequence. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

altering the sequence itself. Interference approaches rely on an sgRNA to guide a catalytically inactive Cas9 (dCas9) to mediate locus-specific transcriptional repression through steric hindrance of biological processes such as blocking RNA polymerase binding/ progression or interfering with the binding of transcription factors [36–39]. Repression strategies extend this concept by fusing a repressive effector domain such as the Krüppel-associated box (KRAB) [36,40], Lys-specific histone demethylase 1 (LSD1) [41], or mSin3 interaction domains [42] to facilitate epigenetic-based repression of the underlying sequence. 2.1. Loss of function approaches with DNA sequence modification 2.1.1. Individual sgRNA targeting Individual sgRNA targeting can be utilized to identify functional non-coding sequences. Given the relatively narrow indel spectrum when using an individual sgRNA, this approach typically requires

additional insight to enhance the pre-experiment probability of successfully identifying functional sequences (Fig. 1a). For example, one study focused on transcription factor binding footprints to identify putative sites to target within the +58 DNase hypersensitive site (DHS) within the enhancer of BCL11A, a known repressor of fetal hemoglobin [43]. Specifically, this study identified 8 potential transcription factor binding sites within the 162 base pair DHS [43,44]. This study used ZFNs and TALENs to target all 8 candidate motifs and identified enhancer function in 1/8 sites. Using a functional genomic footprinting approach for analysis at nucleotide resolution, the region of enhancer function was further reduced to a GATA motif [43]. While this study utilized ZFNs and TALENs to engender DSBs, this strategy could be utilized to inform similar CRISPR-based targeting. Another example involved utilizing patient samples to assist with sgRNA targeting. Overlapping mutations within multiple patient samples were identified in an intergenic region 7.5 kilobases from the transcriptional start site of

M.C. Canver et al. / Methods 121–122 (2017) 118–129

TAL1 that led to the creation on an oncogenic superenhancer [45]. These mutations suggested a narrow site to target via genome editing to study the activity of an aberrant super-enhancer element. Finally, other examples have involved targeting CTCF binding sites within insulators separating topologically associated domains (TADs). These studies have demonstrated that disrupting these CTCF binding sites with indels results in aberrant enhancer activity, changes in gene expression, and disruption of genomic topology [46,47]. Individual sgRNA targeting with Cas9 is also possible in vivo [48,49]. Similarly, in vivo targeting using the Cpf1 nuclease has also been demonstrated by both microinjection and ribonucleoprotein electroporation in mice [50,51]. New strategies for using Cpf1 may allow for multiplex targeting using a single construct. This approach has been used for the simultaneous editing of four targets in vitro and three in vivo [52]. Individual sgRNA targeting is useful for a small number of specific, narrow target(s) often informed by epigenetic marks or transcription factor binding motifs. However, it becomes laborious with multiple targets or to evaluate large regions since large deletions represent a low frequency event following individual sgRNA targeting [12,13,33–35]. The recent identification of regulatory elements lacking typical chromatin signatures highlights the potential shortcomings of targeting only a few predicted functional sequences based on incomplete current knowledge and emphasizes the potential benefits of high-throughput approaches for studying non-coding DNA [53]. 2.1.2. Dual sgRNA targeting for targeted genomic deletion, inversion, and translocations The usage of pairs of sgRNAs, or dual sgRNAs, to produce simultaneous DSBs have a large number of applications including induction of targeted genomic deletions, inversions, and more complex genomic rearrangements such as translocations (Fig. 1b) [54]. Two sgRNAs have been shown to reliably engender large genomic deletions in vitro, including greater than 1 megabase [33,55–59]. Deletions have also been engendered in vivo [60–62]. The use of targeted deletion has been used for a wide variety of applications to study non-coding elements. Notably, targeted deletion has been used to delete composite enhancers as well as enhancer substituents to identify functional sequence within large enhancer regions with characteristic epigenetic marks [63–65]. This strategy has also been used to generate genomic inversions with inverted sequences of greater than 1 megabase [33,55,66–69]. For example, inversion of CTCF binding sites induced dysregulation of genomic topology and alteration of enhancer-promoter contacts, which suggest that enhancers containing CTCF binding sites may not display the orientation independence traditionally ascribed to enhancer elements [70]. Similarly, inversions have validated in situ enhancer element orientation independence, which had been originally described using heterologous reporter constructs [9,63]. A dual sgRNA strategy has been successfully applied to in vivo applications, including mice [66,69,71–75], zebrafish [67], and C. elegans [76]. One study utilized a dual sgRNA approach to engender a 1.2 kilobase deletion and inversion of an insulator regulatory element 12 kilobases upstream of the murine tyrosinase (Tyr) gene, an enzyme necessary for melanin synthesis. CRISPR/Cas9 reagents were delivered through microinjection into fertilized mouse eggs. Using coat pigmentation as a readout, this study confirmed the functionality of this regulatory element as both deletion and inversion of the element resulted in a reduction in both Tyr expression and coat pigmentation [69]. Another study deleted individual DHS and a composite enhancer in mouse embryonic stem cells to ultimately generate germ-line transmitting chimeric mice to examine the role of the enhancer and its substituent components during development [77]. Another notable study disrupted bound-

121

ary elements through deletions, inversions, and duplications to evaluate the effect on embryologic development. The resulting changes in enhancer-promoter interactions resulted in limb malformations [78]. Lastly, utilization of two sgRNA at different chromosomal loci can result in translocations both in vitro and in vivo [79–82], which can allow for chromosomal rearrangements resulting in aberrant enhancer/super-enhancer recruitment [83,84]. The dual sgRNA approach for targeted deletion is useful for interrogation of large regions at low resolution. It is also useful for creating precise deletion events, although indels may be observed at deletion junctions [33]. Dual sgRNA approaches are well suited for generating complex genomic rearrangements such as inversions and translocations. However, the dual sgRNA approach for targeted deletions, inversions, and translocations has been hampered by low throughput, efficiency, and resolution. Recent strategies have been developed for the cloning of pooled libraries consisting of dual sgRNA constructs to allow for pooled deletion/inversion screens [85,86]. This strategy was applied to 51 long non-coding RNAs (lncRNAs) to identify lncRNAs with roles in cancer cell proliferation [87]. These screens may serve as a reliable methodology to increase the throughput of targeted deletion strategies. However, a challenge may be that alternative repair outcomes, such as small indels at each of the dual sgRNA cleavage sites, may limit signal strength in pooled screens, where enrichment phenotypes may be easier to detect than dropout phenotypes. 2.1.3. Tiling sgRNAs/Saturating mutagenesis Pooled CRISPR screening has become widely available and allows for high-throughput and/or multi-loci studies [88–90]. In general, most pooled CRISPR screens have focused on gene knockout for in vitro and in vivo applications [86,91–113]. However, the number of pooled screens targeting non-coding DNA have increased [63,87,114–117]. Notably, a recently developed technique of tiling sgRNAs (‘‘saturating mutagenesis”) has provided a methodology to circumvent the low-throughput, efficiency, and resolution of targeted deletion to interrogate regulatory DNA (Fig. 1c). This approach relies on the same indel spectrum of 1–10 base pairs when utilizing an individual sgRNA [12,13,33–35]. Saturating mutagenesis becomes high-throughput and highresolution by using every sgRNA within a given region in a pooled screening format, which allows for saturating a region with indels to identify functional sequence. The resolution is a function of the number of sgRNA available within a given region. This methodology led to functional fine-mapping of 3.9 kilobases of DHS sequence within the 12 kilobase BCL11A enhancer using a 533 sgRNA library [63]. This study utilized distances between adjacent genomic cleavages between sgRNA within the library as a measure of degree of saturation. For the enhancer-targeted library, there was a median of 4 base pairs between adjacent genomic cleavages and 18 base pairs at 90th percentile. Another study utilized 18,315 sgRNA to mutagenize 715 kilobases of noncoding sequence adjacent to three genes to identify regions involved in resistance to pharmacologic inhibitors by disrupting transcription factor binding and long-range interactions [115]. Yet another group performed saturating mutagenesis of 174 cis regulatory elements including DHS, CTCF sites, and predicted enhancers within a one megabase TAD containing the gene POUF51. This work led to the identification of so-called temp enhancers, whereby the phenotypic effect of enhancer disruption is lost after several rounds of cell division [116]. A final study tiled sgRNA across binding motifs. Specifically, the study utilized 1,116 sgRNA to target p53 binding motifs within 685 genomic loci to identify enhancers required for oncogene-induced senescence. The same study also used 97 sgRNA tiling binding motifs within 73 enhancers for ERa, an estrogenactivated transcription factor implicated in breast cancer. This

122

M.C. Canver et al. / Methods 121–122 (2017) 118–129

approach allowed for identification of novel enhancer elements within both regions [118]. Lastly, a similar strategy was used to tile sgRNAs across regulatory elements with delivery of library sgRNA via HDR to investigate four loci tagged by a fluorescent marker to aide detection of indel perturbations [53]. Enhanced resolution of tiling sgRNA/saturating mutagenesis can be achieved by design of a CRISPR library based on choosing the optimal PAM sequence (and corresponding CRISPR nuclease) for the sequence under study (eg, a GC-rich PAM for a GC-rich region or an AT-rich PAM for an AT-rich region). Alternatively, resolution can be enhanced by using multiple nucleases with unique PAM sequences in combination for the same region [119]. One notable study utilized the S. pyogenes NGG-restricted Cas9 and the S. pyogenes variant NGA-restricted Cas9 to perform saturating mutagenesis of all DHS within the HBS1L-MYB intergenic region [119]. In this study, the usage of NGG-restricted sgRNA alone (n = 2,166 sgRNA) led to a median and 90th percentile gap between adjacent genomic cleavages of 5 and 22.5 base pairs, respectively. The usage of NGA-restricted sgRNA targeting the same region (n = 2,564 sgRNA) led to a median and 90th percentile between adjacent genomic cleavages of 6 and 18 base pairs, respectively. The NGG- and NGA-restricted combination library led to the highest resolution with a total of 4,690 sgRNA resulting in a median of 3 base pairs and 90th percentile of 11 base pairs between adjacent genomic cleavages [119]. With the continual identification of novel Cas9 species and other nucleases with unique PAM sequences [12,13,21–29], the ability to enhance saturating mutagenesis resolution with the combination of nucleases will also continue to improve. Lastly, recent work has also highlighted the importance of taking variants into account during saturating mutagenesis library design for genetically implicated regions marked by common genetic variation. These variant aware libraries help minimize attenuation of sgRNA activity by variants in the cells used for study as well as aide in the detection of single nucleotide polymorphism (SNP)-marked functional sequence [119]. Tiling sgRNA/saturating mutagenesis provides a highthroughput, high resolution methodology to functionally dissect non-coding DNA. This technique has been validated for in vitro applications including large regions and multi-loci studies. While most published studies have utilized NGG-restricted Cas9 [63,114–116], recent work using the NGA-restricted Cas9 variant and Cpf1 have demonstrated that other species can be used for pooled CRISPR screening [119,120]. One potential pitfall encountered with utilizing all sgRNA within a given region is the potential for inclusion of guides with significant off-target potential, which can confound results. To this end, multiple studies have identified potential false positive hits due to targeting of repetitive DNA [119,121,122] including an example when performing saturating mutagenesis of non-coding DNA sequences [119]. Therefore, when performing saturating mutagenesis, it is essential to perform off-target analysis for all sgRNA in the library especially in regions involving repetitive sequences. In addition to off-target analysis, recent work has demonstrated that switching to a modified/adapted tracrRNA sequence can also improve the reliability of data obtained from dropout CRISPR screens and increases the efficacy of library sgRNA [40,123].

(dCas9-KRAB) derived from Kox1 resulted in 5–15 fold gene repression with repression correlating with the levels of sgRNA expression (Fig. 1d). Notably, this approach also was successfully used to repress transcription via enhancer targeting [36]. While the dCas9-only effect was improved through modification of the sgRNA stem loop and hairpin structures, it still resulted in inferior gene repression as compared to the dCas9-KRAB fusion [40]. Another example of using dCas9-KRAB to interrogate noncoding DNA involved targeting the HS2 enhancer within the locus control region (LCR) of the b-globin locus. This led to a reduction in gene expression specific to globin genes and induction of H3K9me3 at HS2 without affecting the other four DHS within the LCR. There was also a decrease in enhancer/promoter accessibility as determined by DNase-sequencing at both HS2 as well as several globin gene promoters [124]. A similar approach was utilized to identify novel distal enhancer elements using a 98,000 sgRNA library of all sgRNA within a 1.29 megabase region flanking the GATA1 and MYC genes, which led to the identification of nine distal enhancer elements [125]. Finally, dCas9-KRAB has been used for genome wide screens for gene targets and lncRNAs, thus demonstrating proof-of-principle for high-throughput, large-scale repression screens [126,127]. Given the role of the histone demethylase LSD1 in enhancer repression, a dCas9-LSD1 fusion was targeted to a known enhancer of Oct4, a regulator of embryonic stem cell state, which resulted in downregulation of Oct4 expression (Fig. 1d). The dCas9-LSD1 fusion was then used to evaluate eight candidate enhancers for a role in regulating embryonic stem cell state ultimately leading to validation of four sites. One validated enhancer region was 10 kilobases upstream of Tbx3, a gene with a known role in the maintenance of pluripotency. Targeting the enhancer region with dCas9-LSD1 resulted in a reduction of Tbx3 at both the RNA and protein level. Notably, there were no changes in enhancerpromoter interaction frequencies. However, there was a 33- to 54-fold reduction in H3K27ac and a 6–8 fold reduction of H3K4me2 at the sgRNA binding site without any associated changes in repressive marks including H3K27me3 and H3K9me3. Surprisingly, upon investigation of the Tbx3 promoter, there was a reduction of H3K27Ac and elevation of H3K27me3 and H3K9me3 when targeting the enhancer with dCas9-KRAB; however, the effect at the promoter was not observed using dCas9LSD1 [41]. These non-indel forming approaches can be used for similar applications as the indel-forming approaches including individual targeting and large-scale screening. However, CRISPRi techniques may be advantageous for investigations where disruption of an enhancer or other regulatory DNA may be cell lethal. The resolution of using dCas9-KRAB or dCas9-LSD1 for saturating mutagenesis as compared to catalytically active Cas9 has yet to be explored. Another notable potential drawback to the CRISPRi approach is that the effect may be transient such that the chromatin may revert to its original state after the CRISPRi machinery is no longer present. However, recent reports have suggested that at least in certain cases the CRISPRi effect may be extremely durable, lasting for weeks to months, although the duration of epigenetic memory may depend on a number of variables including the specific locus, cellular context, and combinations of repressors [128].

2.2. Loss of function approaches without DNA sequence modification 3. Gain of function strategies The initial strategy for studying non-coding DNA without disrupting the underlying sequences relied on CRISPR interference (CRISPRi) to sterically hinder the binding of transcription factors or interference with transcriptional initiation or elongation [36–39]. This strategy led to an 2-fold reduction in gene expression [37]. This technique was enhanced by dCas9 fusion to chromatin modifiers [36,40]. In one study, dCas9 fusion with KRAB domain

Similar to CRISPR methods for loss-of function study, CRISPR based gain-of-function studies focus on two different strategies that are distinguished by the requirement to modify the underlying sequences. The first exploits the endogenous HDR pathway to stimulate repair from an extrachromosomal template to insert customized sequence into the genome. The second strategy relies on

M.C. Canver et al. / Methods 121–122 (2017) 118–129

activation of the underlying sequence without direct modification. This approach relies on dCas9 coupled to transcriptional activation domains [129,130]. 3.1. Gain of function approaches with DNA sequence modification HDR has been an extremely useful tool to study sequence function, but is hampered by low efficiency [131–133]. However, induction of DSBs has been shown to greatly stimulate HDR [134,135]. Inducing HDR is further complicated by cell-cycle regulation such that HDR only occurs during the G2 and late S phase while absent during the G1 phase [136–139]. A variety of strategies have been developed to increase the efficiency of HDR, which have included inhibition of NHEJ machinery, cell cycle synchronization, and optimization of CRISPR and repair template reagents [136,137,139–146]. Despite these technical challenges, HDR has been used for a variety of applications for non-coding DNA including the insertion/knock-in of variants, large fragments of DNA, and reporters at endogenous loci (Fig. 1e) [147–150]. One notable study used single cell labeling of proteins via HDR within neuronal progenitors to allow for imaging of endogenous proteins [151]. HDR has also been successful in vivo for the creation of customized knock-in mice [49,152–155]. Recent work demonstrated 50% of successful HDR alleles to correct a pathogenic SNP using delivery of CRISPR/Cas9 as ribonucleoproteins and adeno-associated virus (AAV) delivery of the repair template. Furthermore, this approach was able to achieve HDR-mediated SNP correction in repopulating hematopoietic stem cells, a quiescent cell type traditionally resistant to HDR [146]. In a similar fashion, one study created four different SNP knock-in mice to examine the effect of these SNPs that had previously been implicated in human infertility [156] while another study created knock-in of two disease risk-associated SNPs implicated for lipid levels both in vitro and in vivo [157]. Finally, HDR approaches have also been used to engender genomic deletions as an alternative to dual sgRNA with NHEJ repair as discussed above. In one study, a rex site was deleted at a TAD boundary through HDR replacement with a NcoI restriction site, which facilitated identification of the desired mutation [158]. One longstanding technical challenge has been identification of causal SNPs and their target genes that have been implicated by genome-wide association studies in human traits and disease [159]. One study proposed a four step pipeline for the identification of causal SNPs through (1) fine-mapping of putative causal variants in conjunction with (2) epigenetic profiling. The study identified rs339331 as a candidate causal SNP that was suggested to be located within a regulatory element controlling RFX6 expression. Then the authors used (3) epigenetic editing to demonstrate that targeting the rs339331 region with TALE-LSD1 fusion reduced RFX6 expression, a finding consistent with enhancer function as discussed above. Finally, rs339331 was inserted into cell lines via (4) TALE-mediated HDR to evaluate its functional consequences [160]. This pipeline provides a systematic approach for the evaluation of genetically implicated variants marking regulatory DNA. Recent reports have suggested an alternative approach to the stimulated HDR following nuclease cleavage paradigm to insert/ alter sequence. Specifically, two studies fused cytidine deaminase enzymes to dCas9 for site-specific C to T or G to A substitutions in an 5 base pair window in the absence of nuclease-mediated cleavage [161,162]. To further this technique, other groups used a variant dCas9-AID fusion that was capable of creating C and G substitutions to the three other bases at appreciable frequencies (Fig. 1f). This system was used to identify previously reported and novel mutations associated with resistance to the tyrosine kinase inhibitor imatinib [163]. Finally, another group used dCas9-AID with its nuclear export signal deleted in conjunction

123

with an sgRNA with two MS2 activation domains to generate significant levels of genomic diversity to drive directed evolution [164]. Saturation mutagenesis, distinct from the tiling sgRNA/saturating mutagenesis discussed above, is an HDR-based technique for comprehensive mutagenesis of narrow regions. To be explicit, we are using the term ‘‘saturating mutagenesis” to describe design of all possible sgRNAs for a given nuclease within a given target sequence, and the term ‘‘saturation mutagenesis” to describe introduction of all possible sequence variants via HDR within a given target sequence. Saturation mutagenesis has been used for a 6 base pair region within exon 18 of the BRCA1 gene to generate all possible single nucleotide variants via HDR from a library of single nucleotide substituted templates to examine the effects of each substitution on transcript levels [165]. This approach provides a comprehensive interrogation of a given region. While this approach has only been demonstrated in coding sequence, this could be used for non-coding regions such as for the evaluation of transcription factor binding motifs. While this strategy benefits from its comprehensive approach to a region, it is limited in its throughput and requirement for narrow regions. Lastly, another recent report demonstrated the use of HDR to integrate a library of 12,000 175 base pair donors with 100 base pair variable regions into heterochromatin to confirm computationally predicted sequences associated with chromatin accessibility in situ [166]. Several studies have attempted to circumvent the HDR requirement for the insertion of exogenous sequence into the genome. One study achieved sequence knock-in of a fluorescent marker from a repair template by exploiting microhomology-mediated end joining (MMEJ), a repair process present during G1 and early S phases that relies on nearby microhomologous sequences of 2– 25 bp in length (Fig. 1e). This approach was also successful in vivo in both silkworms and frogs [167]. It had been previously shown that simultaneous ZFN-mediated cleavage of both a genomic locus and a plasmid containing the same cleavage recognition site led to integration of the plasmid (Fig. 1g) [168]. Using a similar strategy, others demonstrated homology-independent integration in zebrafish with simultaneous CRISPR/Cas9-mediated DSB induction at a genomic locus with concurrent plasmid linearization [169,170]. Finally, recent work showed that an NHEJ-mediated targeted integration approach involving concurrent genomic locus and plasmid cleavage outperformed HDR and MMEJ integration, which could be used in both dividing and non-dividing cells. The authors also demonstrated its use in vivo [171]. 3.2. Gain of function approaches without DNA sequence modification The initial strategies for gain-of-function study of non-coding DNA without modifying the underlying sequences have relied on CRISPR activation (CRISPRa) to augment transcription. One initial study created a fusion of VP64 activation domain to the Cterminus of Cas9, which led to transcriptional activation (Fig. 1h). In this study, the 3’ end of the sgRNA was also modified with two copies of MS2 bacteriophage coat protein-binding aptamers to allow for sgRNA recruitment of activation domains [130]. Taken together, these two modifications allowed for CRISPR-mediated gene activation. Other studies have used alternative activation domains such as VP160 [172]. This approach was furthered in a subsequent study that modified Cas9 to include MS2 protein-interacting RNA aptamers on the sgRNA’s tetraloop and stem loop 2, which outperformed dCas9-VP64 fusion 3–5 fold individually and a 12-fold increase when both aptamers were included. The combination of these modifications with dCas9-VP64 further increased activation of gene expression [173]. Notably, the MS2 aptamer modification on the tetraloop and stem loop 2 was more effective than the 3’

124

M.C. Canver et al. / Methods 121–122 (2017) 118–129

end fusion of MS2 [130,173]. In this same study, dCas9 was modified to add additional activation domains including VP64, the NFjB trans-activating subunit p65, and the activation domain from human heat shock factor 1 (HSF1). Collectively, the optimal system for transcriptional activation was demonstrated to include sgRNAs with MS2 aptamer modifications on the tetraloop and stem loop 2, dCas9-VP64, and the addition of activators MS2-p65-HSF1. This system was successfully used at both individual loci and for genome scale activation screening [173,174]. A recent study applied the CRISPRa strategy to non-coding regions by using the dCas9-VP64 system for an unbiased screen to identify functional enhancer sequences by tiling sgRNA across a >100 kilobase genomic region genetically implicated to play a role in autoimmunity. This approach led to the identification of multiple enhancer elements [175]. Finally, another approach to transcriptional activation is the fusion of dCas9 to the catalytic core of the human acetyltransferase p300 to allow for the catalysis of H3K27Ac at target loci (Fig. 1h). dCas9-p300 outperformed dCas9-VP64 to increase gene expression when targeting multiple promoters. Notably, targeting of the HS2 enhancer within the bglobin locus by dCas9-p300 led to increased expression of globin genes at levels higher than dCas9-VP64. The authors further showed that activation correlated with increased H3K27Ac at both promoters and enhancer regions [176]. 4. General considerations There are many general considerations that are applicable to both loss-of-function and gain-of-function approaches to studying non-coding DNA. While these considerations are beyond the scope of this review, we will briefly highlight several important issues. For example, several studies have investigated methods to predict/enhance sgRNA activity at both the individual sgRNA level [94,98,177–188] and for large-scale pooled screening [94,189]. In addition, many methods have been developed to predict/detect off-target cleavages [35,178,190–193]. Other methods to minimize off-targets have been developed including the use of truncated sgRNAs, dimeric RNA-guided FokI nucleases, and high-fidelity Cas9 mutants [34,54,194–197]. Recent work has provided further insight to the targeting specificity of Cpf1 [120,198,199]. It is important to maximize sgRNA activity and minimize off-target cleavages as allowed by the experimental design. Methodologies continue to emerge that allow for increased spatiotemporal control of genome editing experiments, such as chemical- and light-inducible Cas9 [200–202] as well as identification of naturally occurring inhibitors of Cas9 that may be exploited as an on/off switch [203]. Furthermore, combining CRISPR pooled screens with RNA-sequencing at the single cell level will aid both loss-of-function and gain-of-function experiments [204–206]. Finally, systems for simultaneous repression and activation of gene expression through the use of scaffold RNAs may usher in an era where combinatorial loss- and gain-of-function studies are possible to dissect complex transcriptional networks [207]. 5. Conclusions The study of regulatory DNA has been plagued by a paucity of methods to interrogate non-coding sequences in their native chromatin, chromosomal, and cellular context. CRISPR-based technologies continue to emerge at a rapid pace, which offer new methodologies to study non-coding DNA for laboratory experiments and therapeutic applications [43,63,208–210]. CRISPRbased strategies allow for efficient loss- and gain-of-function studies, which includes approaches with and without modification of the targeted DNA sequences. Advancements have continued to

improve throughput, efficiency, and specificity of CRISPR-based methodologies. CRISPR technologies will continue to unmask and further the understanding of the non-coding genome. Author contributions M.C.C. reviewed the literature. M.C.C., D.E.B., and S.H.O. wrote the manuscript. Conflict of interest disclosures M.C.C, D.E.B., and S.H.O. declare no competing financial interests. Acknowledgments M.C.C. is supported by a National Institute of Diabetes and Digestive and Kidney Diseases (NIDDK) Award (F30DK10335901A1). D.E.B. is supported by NIDDK (K08DK093705, R03DK109232), NHLBI (DP2OD022716), Burroughs Wellcome Fund, Doris Duke Charitable Foundation Innovations in Clinical Research Award, ASH Scholar Award, Charles H. Hood Foundation Child Health Research Award, and Cooley’s Anemia Foundation fellowship. S.H.O. is supported by an award from the NHLBI award (P01HL032262) and an award from the NIDDK (P30DK049216, Center of Excellence in Molecular Hematology). References [1] E.S. Lander, L.M. Linton, B. Birren, C. Nusbaum, M.C. Zody, J. Baldwin, K. Devon, K. Dewar, M. Doyle, W. FitzHugh, R. Funke, D. Gage, K. Harris, A. Heaford, J. Howland, L. Kann, J. Lehoczky, R. LeVine, P. McEwan, K. McKernan, J. Meldrim, J.P. Mesirov, C. Miranda, W. Morris, J. Naylor, C. Raymond, M. Rosetti, R. Santos, A. Sheridan, C. Sougnez, N. Stange-Thomann, N. Stojanovic, A. Subramanian, D. Wyman, J. Rogers, J. Sulston, R. Ainscough, S. Beck, D. Bentley, J. Burton, C. Clee, N. Carter, A. Coulson, R. Deadman, P. Deloukas, A. Dunham, I. Dunham, R. Durbin, L. French, D. Grafham, S. Gregory, T. Hubbard, S. Humphray, A. Hunt, M. Jones, C. Lloyd, A. McMurray, L. Matthews, S. Mercer, S. Milne, J.C. Mullikin, A. Mungall, R. Plumb, M. Ross, R. Shownkeen, S. Sims, R.H. Waterston, R.K. Wilson, L.W. Hillier, J.D. McPherson, M.A. Marra, E. R. Mardis, L.A. Fulton, A.T. Chinwalla, K.H. Pepin, W.R. Gish, S.L. Chissoe, M.C. Wendl, K.D. Delehaunty, T.L. Miner, A. Delehaunty, J.B. Kramer, L.L. Cook, R.S. Fulton, D.L. Johnson, P.J. Minx, S.W. Clifton, T. Hawkins, E. Branscomb, P. Predki, P. Richardson, S. Wenning, T. Slezak, N. Doggett, J.F. Cheng, A. Olsen, S. Lucas, C. Elkin, E. Uberbacher, M. Frazier, R.A. Gibbs, D.M. Muzny, S.E. Scherer, J.B. Bouck, E.J. Sodergren, K.C. Worley, C.M. Rives, J.H. Gorrell, M.L. Metzker, S. L. Naylor, R.S. Kucherlapati, D.L. Nelson, G.M. Weinstock, Y. Sakaki, A. Fujiyama, M. Hattori, T. Yada, A. Toyoda, T. Itoh, C. Kawagoe, H. Watanabe, Y. Totoki, T. Taylor, J. Weissenbach, R. Heilig, W. Saurin, F. Artiguenave, P. Brottier, T. Bruls, E. Pelletier, C. Robert, P. Wincker, D.R. Smith, L. DoucetteStamm, M. Rubenfield, K. Weinstock, H.M. Lee, J. Dubois, A. Rosenthal, M. Platzer, G. Nyakatura, S. Taudien, A. Rump, H. Yang, J. Yu, J. Wang, G. Huang, J. Gu, L. Hood, L. Rowen, A. Madan, S. Qin, R.W. Davis, N.A. Federspiel, A.P. Abola, M.J. Proctor, R.M. Myers, J. Schmutz, M. Dickson, J. Grimwood, D.R. Cox, M.V. Olson, R. Kaul, C. Raymond, N. Shimizu, K. Kawasaki, S. Minoshima, G.A. Evans, M. Athanasiou, R. Schultz, B.A. Roe, F. Chen, H. Pan, J. Ramser, H. Lehrach, R. Reinhardt, W.R. McCombie, M. de la Bastide, N. Dedhia, H. Blocker, K. Hornischer, G. Nordsiek, R. Agarwala, L. Aravind, J.A. Bailey, A. Bateman, S. Batzoglou, E. Birney, P. Bork, D.G. Brown, C.B. Burge, L. Cerutti, H.C. Chen, D. Church, M. Clamp, R.R. Copley, T. Doerks, S.R. Eddy, E.E. Eichler, T.S. Furey, J. Galagan, J.G. Gilbert, C. Harmon, Y. Hayashizaki, D. Haussler, H. Hermjakob, K. Hokamp, W. Jang, L.S. Johnson, T.A. Jones, S. Kasif, A. Kaspryzk, S. Kennedy, W. J. Kent, P. Kitts, E.V. Koonin, I. Korf, D. Kulp, D. Lancet, T.M. Lowe, A. McLysaght, T. Mikkelsen, J.V. Moran, N. Mulder, V.J. Pollara, C.P. Ponting, G. Schuler, J. Schultz, G. Slater, A.F. Smit, E. Stupka, J. Szustakowski, D. ThierryMieg, J. Thierry-Mieg, L. Wagner, J. Wallis, R. Wheeler, A. Williams, Y.I. Wolf, K.H. Wolfe, S.P. Yang, R.F. Yeh, F. Collins, M.S. Guyer, J. Peterson, A. Felsenfeld, K.A. Wetterstrand, A. Patrinos, M.J. Morgan, P. de Jong, J.J. Catanese, K. Osoegawa, H. Shizuya, S. Choi, Y.J. Chen, International human genome sequencing, initial sequencing and analysis of the human genome, Nature 409 (2001) 860–921. [2] M. Kellis, B. Wold, M.P. Snyder, B.E. Bernstein, A. Kundaje, G.K. Marinov, L.D. Ward, E. Birney, G.E. Crawford, J. Dekker, I. Dunham, L.L. Elnitski, P.J. Farnham, E.A. Feingold, M. Gerstein, M.C. Giddings, D.M. Gilbert, T.R. Gingeras, E.D. Green, R. Guigo, T. Hubbard, J. Kent, J.D. Lieb, R.M. Myers, M.J. Pazin, B. Ren, J. A. Stamatoyannopoulos, Z. Weng, K.P. White, R.C. Hardison, Defining

M.C. Canver et al. / Methods 121–122 (2017) 118–129 functional DNA elements in the human genome, Proc. Natl. Acad. Sci. U. S. A. 111 (2014) 6131–6138. [3] K. Chi, The dark side of the human genome, Nature 538 (2016) 275–277. [4] I. Dunham, A. Kundaje, S.F. Aldred, P.J. Collins, C.A. Davis, F. Doyle, C.B. Epstein, S. Frietze, J. Harrow, R. Kaul, J. Khatun, B.R. Lajoie, S.G. Landt, B.-K. Lee, F. Pauli, K.R. Rosenbloom, P. Sabo, A. Safi, A. Sanyal, N. Shoresh, J.M. Simon, L. Song, N.D. Trinklein, R.C. Altshuler, E. Birney, J.B. Brown, C. Cheng, S. Djebali, X. Dong, I. Dunham, J. Ernst, T.S. Furey, M. Gerstein, B. Giardine, M. Greven, R.C. Hardison, R.S. Harris, J. Herrero, M.M. Hoffman, S. Iyer, M. Kellis, J. Khatun, P. Kheradpour, A. Kundaje, T. Lassmann, Q. Li, X. Lin, G.K. Marinov, A. Merkel, A. Mortazavi, S.C.J. Parker, T.E. Reddy, J. Rozowsky, F. Schlesinger, R.E. Thurman, J. Wang, L.D. Ward, T.W. Whitfield, S.P. Wilder, W. Wu, H.S. Xi, K.Y. Yip, J. Zhuang, B.E. Bernstein, E. Birney, I. Dunham, E.D. Green, C. Gunter, M. Snyder, M.J. Pazin, R.F. Lowdon, L.A.L. Dillon, L.B. Adams, C.J. Kelly, J. Zhang, J. R. Wexler, E.D. Green, P.J. Good, E.A. Feingold, B.E. Bernstein, E. Birney, G.E. Crawford, J. Dekker, L. Elnitski, P.J. Farnham, M. Gerstein, M.C. Giddings, T.R. Gingeras, E.D. Green, R. Guigó, R.C. Hardison, T.J. Hubbard, M. Kellis, W.J. Kent, J.D. Lieb, E.H. Margulies, R.M. Myers, M. Snyder, J.A. Stamatoyannopoulos, S.A. Tenenbaum, Z. Weng, K.P. White, B. Wold, J. Khatun, Y. Yu, J. Wrobel, B.A. Risk, H.P. Gunawardena, H.C. Kuiper, C.W. Maier, L. Xie, X. Chen, M.C. Giddings, B.E. Bernstein, C.B. Epstein, N. Shoresh, J. Ernst, P. Kheradpour, T.S. Mikkelsen, S. Gillespie, A. Goren, O. Ram, X. Zhang, L. Wang, R. Issner, M.J. Coyne, T. Durham, M. Ku, T. Truong, L.D. Ward, R.C. Altshuler, M.L. Eaton, M. Kellis, S. Djebali, C.A. Davis, A. Merkel, A. Dobin, T. Lassmann, A. Mortazavi, A. Tanzer, J. Lagarde, W. Lin, F. Schlesinger, C. Xue, G.K. Marinov, J. Khatun, B.A. Williams, C. Zaleski, J. Rozowsky, M. Röder, F. Kokocinski, R.F. Abdelhamid, T. Alioto, I. Antoshechkin, M.T. Baer, P. Batut, I. Bell, K. Bell, S. Chakrabortty, X. Chen, J. Chrast, J. Curado, T. Derrien, J. Drenkow, E. Dumais, J. Dumais, R. Duttagupta, M. Fastuca, K. Fejes-Toth, P. Ferreira, S. Foissac, M.J. Fullwood, H. Gao, D. Gonzalez, A. Gordon, H.P. Gunawardena, C. Howald, S. Jha, R. Johnson, P. Kapranov, B. King, C. Kingswood, G. Li, O.J. Luo, E. Park, J.B. Preall, K. Presaud, P. Ribeca, B.A. Risk, D. Robyr, X. Ruan, M. Sammeth, K.S. Sandhu, L. Schaeffer, L.-H. See, A. Shahab, J. Skancke, A.M. Suzuki, H. Takahashi, H. Tilgner, D. Trout, N. Walters, H. Wang, J. Wrobel, Y. Yu, Y. Hayashizaki, J. Harrow, M. Gerstein, T. J. Hubbard, A. Reymond, S.E. Antonarakis, G.J. Hannon, M.C. Giddings, Y. Ruan, B. Wold, P. Carninci, R. Guigó, T.R. Gingeras, K.R. Rosenbloom, C.A. Sloan, K. Learned, V.S. Malladi, M.C. Wong, G.P. Barber, M.S. Cline, T.R. Dreszer, S.G. Heitner, D. Karolchik, W.J. Kent, V.M. Kirkup, L.R. Meyer, J.C. Long, M. Maddren, B.J. Raney, T.S. Furey, L. Song, L.L. Grasfeder, P.G. Giresi, B.-K. Lee, A. Battenhouse, N.C. Sheffield, J.M. Simon, K.A. Showers, A. Safi, D. London, A.A. Bhinge, C. Shestak, M.R. Schaner, S. Ki Kim, Z.Z. Zhang, P.A. Mieczkowski, J.O. Mieczkowska, Z. Liu, R.M. McDaniell, Y. Ni, N.U. Rashid, M.J. Kim, S. Adar, Z. Zhang, T. Wang, D. Winter, D. Keefe, E. Birney, V.R. Iyer, J.D. Lieb, G.E. Crawford, G. Li, K.S. Sandhu, M. Zheng, P. Wang, O.J. Luo, A. Shahab, M.J. Fullwood, X. Ruan, Y. Ruan, R.M. Myers, F. Pauli, B.A. Williams, J. Gertz, G.K. Marinov, T.E. Reddy, J. Vielmetter, E. Partridge, D. Trout, K.E. Varley, C. Gasper, A. Bansal, S. Pepke, P. Jain, H. Amrhein, K.M. Bowling, M. Anaya, M.K. Cross, B. King, M.A. Muratet, I. Antoshechkin, K.M. Newberry, K. McCue, A.S. Nesmith, K.I. Fisher-Aylor, B. Pusey, G. DeSalvo, S.L. Parker, S. Balasubramanian, N.S. Davis, S.K. Meadows, T. Eggleston, C. Gunter, J.S. Newberry, S.E. Levy, D.M. Absher, A. Mortazavi, W.H. Wong, B. Wold, M.J. Blow, A. Visel, L.A. Pennachio, L. Elnitski, E.H. Margulies, S.C.J. Parker, H.M. Petrykowska, A. Abyzov, B. Aken, D. Barrell, G. Barson, A. Berry, A. Bignell, V. Boychenko, G. Bussotti, J. Chrast, C. Davidson, T. Derrien, G. Despacio-Reyes, M. Diekhans, I. Ezkurdia, A. Frankish, J. Gilbert, J.M. Gonzalez, E. Griffiths, R. Harte, D.A. Hendrix, C. Howald, T. Hunt, I. Jungreis, M. Kay, E. Khurana, F. Kokocinski, J. Leng, M.F. Lin, J. Loveland, Z. Lu, D. Manthravadi, M. Mariotti, J. Mudge, G. Mukherjee, C. Notredame, B. Pei, J.M. Rodriguez, G. Saunders, A. Sboner, S. Searle, C. Sisu, C. Snow, C. Steward, A. Tanzer, E. Tapanari, M.L. Tress, M.J. van Baren, N. Walters, S. Washietl, L. Wilming, A. Zadissa, Z. Zhang, M. Brent, D. Haussler, M. Kellis, A. Valencia, M. Gerstein, A. Reymond, R. Guigó, J. Harrow, T.J. Hubbard, S.G. Landt, S. Frietze, A. Abyzov, N. Addleman, R.P. Alexander, R.K. Auerbach, S. Balasubramanian, K. Bettinger, N. Bhardwaj, A.P. Boyle, A.R. Cao, P. Cayting, A. Charos, Y. Cheng, C. Cheng, C. Eastman, G. Euskirchen, J.D. Fleming, F. Grubert, L. Habegger, M. Hariharan, A. Harmanci, S. Iyengar, V.X. Jin, K.J. Karczewski, M. Kasowski, P. Lacroute, H. Lam, N. Lamarre-Vincent, J. Leng, J. Lian, M. Lindahl-Allen, R. Min, B. Miotto, H. Monahan, Z. Moqtaderi, X.J. Mu, H. O’Geen, Z. Ouyang, D. Patacsil, B. Pei, D. Raha, L. Ramirez, B. Reed, J. Rozowsky, A. Sboner, M. Shi, C. Sisu, T. Slifer, H. Witt, L. Wu, X. Xu, K.-K. Yan, X. Yang, K.Y. Yip, Z. Zhang, K. Struhl, S.M. Weissman, M. Gerstein, P.J. Farnham, M. Snyder, S.A. Tenenbaum, L.O. Penalva, F. Doyle, S. Karmakar, S.G. Landt, R.R. Bhanvadia, A. Choudhury, M. Domanus, L. Ma, J. Moran, D. Patacsil, T. Slifer, A. Victorsen, X. Yang, M. Snyder, K.P. White, T. Auer, L. Centanin, M. Eichenlaub, F. Gruhl, S. Heermann, B. Hoeckendorf, D. Inoue, T. Kellner, S. Kirchmaier, C. Mueller, R. Reinhardt, L. Schertel, S. Schneider, R. Sinn, B. Wittbrodt, J. Wittbrodt, Z. Weng, T.W. Whitfield, J. Wang, P.J. Collins, S.F. Aldred, N.D. Trinklein, E.C. Partridge, R.M. Myers, J. Dekker, G. Jain, B.R. Lajoie, A. Sanyal, G. Balasundaram, D.L. Bates, R. Byron, T.K. Canfield, M.J. Diegel, D. Dunn, A.K. Ebersol, T. Frum, K. Garg, E. Gist, R.S. Hansen, L. Boatman, E. Haugen, R. Humbert, G. Jain, A.K. Johnson, E.M. Johnson, T.V. Kutyavin, B.R. Lajoie, K. Lee, D. Lotakis, M.T. Maurano, S.J. Neph, F.V. Neri, E.D. Nguyen, H. Qu, A.P. Reynolds, V. Roach, E. Rynes, P. Sabo, M.E. Sanchez, R.S. Sandstrom, A. Sanyal, A.O. Shafer, A.B. Stergachis, S. Thomas, R.E. Thurman, B. Vernot, J. Vierstra, S. Vong, H. Wang, M.A. Weaver, Y. Yan, M. Zhang, J.M. Akey, M. Bender, M.O. Dorschner, M. Groudine, M.J. McCoss, P.

[5] [6]

[7]

[8]

[9] [10] [11] [12]

[13]

[14] [15] [16] [17]

[18] [19] [20]

[21]

[22]

[23]

[24]

[25]

[26]

[27]

[28]

[29]

125

Navas, G. Stamatoyannopoulos, R. Kaul, J. Dekker, J.A. Stamatoyannopoulos, I. Dunham, K. Beal, A. Brazma, P. Flicek, J. Herrero, N. Johnson, D. Keefe, M. Lukk, N.M. Luscombe, D. Sobral, J.M. Vaquerizas, S.P. Wilder, S. Batzoglou, A. Sidow, N. Hussami, S. Kyriazopoulou-Panagiotopoulou, M.W. Libbrecht, M.A. Schaub, A. Kundaje, R.C. Hardison, W. Miller, B. Giardine, R.S. Harris, W. Wu, P.J. Bickel, B. Banfai, N.P. Boley, J.B. Brown, H. Huang, Q. Li, J.J. Li, W.S. Noble, J.A. Bilmes, O.J. Buske, M.M. Hoffman, A.D. Sahu, P.V. Kharchenko, P.J. Park, D. Baker, J. Taylor, Z. Weng, S. Iyer, X. Dong, M. Greven, X. Lin, J. Wang, H.S. Xi, J. Zhuang, M. Gerstein, R.P. Alexander, S. Balasubramanian, C. Cheng, A. Harmanci, L. Lochovsky, R. Min, X.J. Mu, J. Rozowsky, K.-K. Yan, K.Y. Yip, E. Birney, An integrated encyclopedia of DNA elements in the human genome, Nature 489 (2012) 57–74. M. Haeussler, J.S. Joly, When needles look like hay: How to find tissue-specific enhancers in model organism genomes, Dev. Biol. 350 (2011) 239–254. R.P. Patwardhan, J.B. Hiatt, D.M. Witten, M.J. Kim, R.P. Smith, D. May, C. Lee, J. M. Andrie, S.-I. Lee, G.M. Cooper, N. Ahituv, L.A. Pennacchio, J. Shendure, Massively parallel functional dissection of mammalian enhancers in vivo, Nat. Biotechnol. 30 (2012) 265–270. A. Melnikov, A. Murugan, X. Zhang, T. Tesileanu, L. Wang, P. Rogov, S. Feizi, A. Gnirke, C.G. Callan, J.B. Kinney, M. Kellis, E.S. Lander, T.S. Mikkelsen, Systematic dissection and optimization of inducible enhancers in human cells using a massively parallel reporter assay, Nat. Biotechnol. 30 (2012) 271–277. J.C. Ulirsch, S.K. Nandakumar, L. Wang, R. Do, T.S. Mikkelsen, V.G. Sankaran, Systematic functional dissection of common genetic variation affecting red blood cell traits, Cell 165 (2016) 1530–1545. J. Banerji, S. Rusconi, W. Schaffner, Expression of a b-globin gene is enhanced by remote SV40 DNA sequences, Cell 27 (1981) 299–308. J.K. Joung, J.D. Sander, TALENs: a widely applicable technology for targeted genome editing, Nat. Rev. Mol. Cell Biol. 14 (2013) 49–55. F.D. Urnov, E.J. Rebar, M.C. Holmes, H.S. Zhang, P.D. Gregory, Genome editing with engineered zinc finger nucleases, Nat. Rev. Genet. 11 (2010) 636–646. L. Cong, F.A. Ran, D. Cox, S. Lin, R. Barretto, N. Habib, P.D. Hsu, X. Wu, W. Jiang, L.A. Marraffini, F. Zhang, Multiplex genome engineering using CRISPR/Cas systems, Science 339 (2013) 819–823. P. Mali, L. Yang, K.M. Esvelt, J. Aach, M. Guell, J.E. DiCarlo, J.E. Norville, G.M. Church, RNA-guided human genome engineering via Cas9, Science 339 (2013) 823–826. P.D. Hsu, E.S. Lander, F. Zhang, Development and applications of CRISPR-Cas9 for genome engineering, Cell 157 (2014) 1262–1278. J.D. Sander, J.K. Joung, CRISPR-Cas systems for editing, regulating and targeting genomes, Nat. Biotechnol. 32 (2014) 347–355. C.M. Nowak, S. Lawson, M. Zerez, L. Bleris, Guide RNA engineering for versatile Cas9 functionality, Nucleic Acids Res. (2016). X. Wang, X. Huang, X. Fang, Y. Zhang, W. Wang, CRISPR-Cas9 system as a versatile tool for genome engineering in human cells, Mol. Ther. Acids 5 (2016) e388. A.C. Komor, A.H. Badran, D.R. Liu, CRISPR-based technologies for the manipulation of eukaryotic genomes, Cell 1–17 (2017). R. Barrangou, J.A. Doudna, Applications of CRISPR technologies in research and beyond, Nat. Biotechnol. 34 (2016) 933–941. R. Barrangou, C. Fremaux, H. Deveau, M. Richards, P. Boyaval, S. Moineau, D.A. Romero, P. Horvath, CRISPR provides acquired resistance against viruses in prokaryotes, Science 315 (2007) 1709–1712. B.P. Kleinstiver, M.S. Prew, S.Q. Tsai, V.V. Topkar, N.T. Nguyen, Z. Zheng, A.P. Gonzales, Z. Li, R.T. Peterson, J.R. Yeh, M.J. Aryee, J.K. Joung, Engineered CRISPR-Cas9 nucleases with altered PAM specificities, Nature 523 (2015) 481–485. C. Anders, O. Niewoehner, A. Duerst, M. Jinek, Structural basis of PAMdependent target DNA recognition by the Cas9 endonuclease, Nature 513 (2014) 569–573. H. Hirano, J.S. Gootenberg, T. Horii, O.O. Abudayyeh, M. Kimura, P.D. Hsu, T. Nakane, R. Ishitani, I. Hatada, F. Zhang, H. Nishimasu, O. Nureki, Structure and engineering of Francisella novicida Cas9, Cell 164 (2016) 950–961. B.P. Kleinstiver, M.S. Prew, S.Q. Tsai, N.T. Nguyen, V.V. Topkar, Z. Zheng, J.K. Joung, Broadening the targeting range of Staphylococcus aureus CRISPR-Cas9 by modifying PAM recognition, Nat. Biotechnol. 33 (2015) 1293–1298. K.M. Esvelt, P. Mali, J.L. Braff, M. Moosburner, S.J. Yaung, G.M. Church, Orthogonal Cas9 proteins for RNA-guided gene regulation and editing, Nat. Methods 10 (2013) 1116–1121. T. Karvelis, G. Gasiunas, J. Young, G. Bigelyte, A. Silanskas, M. Cigan, V. Siksnys, Rapid characterization of CRISPR-Cas9 protospacer adjacent motif sequence elements, Genome Biol. 16 (2015) 253. F.A. Ran, L. Cong, W.X. Yan, D.A. Scott, J.S. Gootenberg, A.J. Kriz, B. Zetsche, O. Shalem, X. Wu, K.S. Makarova, E.V. Koonin, P.A. Sharp, F. Zhang, In vivo genome editing using Staphylococcus aureus Cas9, Nature 520 (2015) 186– 191. S. Shmakov, O.O. Abudayyeh, K.S. Makarova, Y.I. Wolf, J.S. Gootenberg, E. Semenova, L. Minakhin, J. Joung, S. Konermann, K. Severinov, F. Zhang, E.V. Koonin, Discovery and functional characterization of diverse class 2 CRISPRCas systems, Mol. Cell 60 (2015) 385–397. B. Zetsche, J.S. Gootenberg, O.O. Abudayyeh, A. Regev, E.V. Koonin, F. Zhang, T. Pam, B. Zetsche, J.S. Gootenberg, O.O. Abudayyeh, I.M. Slaymaker, K.S. Makarova, P. Essletzbichler, S.E. Volz, J. Joung, J. van der Oost, A. Regev, E.V.

126

[30]

[31] [32]

[33]

[34]

[35]

[36]

[37]

[38]

[39]

[40]

[41]

[42]

[43]

[44]

[45]

[46]

[47]

[48]

[49]

[50]

[51]

[52]

M.C. Canver et al. / Methods 121–122 (2017) 118–129 Koonin, F. Zhang, Cpf1 is a single RNA-guided endonuclease of a class 2 CRISPR-Cas system, Cell 163 (2015) 759–771. H. Yang, P. Gao, K.R. Rajashankar, D.J. Patel, PAM-dependent target DNA recognition and cleavage by C2c1 CRISPR-Cas endonuclease, Cell 167 (2016). 1814–1828.e12. F.A. Ran, P.D. Hsu, J. Wright, V. Agarwala, D.A. Scott, F. Zhang, Genome engineering using the CRISPR-Cas9 system, Nat. Protoc. 8 (2013) 2281–2308. R. Lopes, G. Korkmaz, R. Agami, Applying CRISPR-Cas9 tools to identify and characterize transcriptional enhancers, Nat. Rev. Mol. Cell Biol. 17 (2016) 597–604. M.C. Canver, D.E. Bauer, A. Dass, Y.Y. Yien, J. Chung, T. Masuda, T. Maeda, B.H. Paw, S.H. Orkin, Characterization of genomic deletion efficiency mediated by clusted regularly interspaced palindromic repeats (CRISPR)/Cas9 nuclease system in mammalian cells, J. Biol. Chem. 289 (2014) 21312–21324. F.A. Ran, P.D. Hsu, C.Y. Lin, J.S. Gootenberg, S. Konermann, A.E. Trevino, D.A. Scott, A. Inoue, S. Matoba, Y. Zhang, F. Zhang, Double nicking by RNA-guided CRISPR Cas9 for enhanced genome editing specificity, Cell 154 (2013) 1380– 1389. P.D. Hsu, D.A. Scott, J.A. Weinstein, F.A. Ran, S. Konermann, V. Agarwala, Y. Li, E.J. Fine, X. Wu, O. Shalem, T.J. Cradick, L.A. Marraffini, G. Bao, F. Zhang, DNA targeting specificity of RNA-guided Cas9 nucleases, Nat. Biotechnol. 31 (2013) 827–832. L.A. Gilbert, M.H. Larson, L. Morsut, Z. Liu, G.A. Brar, S.E. Torres, N. SternGinossar, O. Brandman, E.H. Whitehead, J.A. Doudna, W.A. Lim, J.S. Weissman, L.S. Qi, CRISPR-mediated modular RNA-guided regulation of transcription in eukaryotes, Cell 154 (2013) 442–451. L.S. Qi, M.H. Larson, L.A. Gilbert, J.A. Doudna, J.S. Weissman, A.P. Arkin, W.A. Lim, Repurposing CRISPR as an RNA-guided platform for sequence-specific control of gene expression, Cell 152 (2013) 1173–1183. M.H. Larson, L.A. Gilbert, X. Wang, W.A. Lim, J.S. Weissman, L.S. Qi, CRISPR interference (CRISPRi) for sequence-specific control of gene expression, Nat. Protoc. 8 (2013) 2180–2196. X. Gao, J.C.H. Tsang, F. Gaba, D. Wu, L. Lu, P. Liu, Comparison of TALE designer transcription factors and the CRISPR/dCas9 in regulation of gene expression by targeting enhancers, Nucleic Acids Res. 42 (2014). B. Chen, L. a Gilbert, B. a Cimini, J. Schnitzbauer, W. Zhang, G.-W. Li, J. Park, E. H. Blackburn, J.S. Weissman, L.S. Qi, B. Huang, Dynamic imaging of genomic loci in living human cells by an optimized CRISPR/Cas system, Cell 155 (2013) 1479–1491. N.A. Kearns, H. Pham, B. Tabak, R.M. Genga, N.J. Silverstein, M. Garber, R. Maehr, Functional annotation of native enhancers with a Cas9-histone demethylase fusion, Nat. Methods 12 (2015) 401–403. S. Konermann, M.D. Brigham, A.E. Trevino, P.D. Hsu, M. Heidenreich, L. Cong, R.J. Platt, D.A. Scott, G.M. Church, F. Zhang, Optical control of mammalian endogenous transcription and epigenetic states, Nature 500 (2013) 472–476. J. Vierstra, A. Reik, K.-H. Chang, S. Stehling-Sun, Y. Zhou, S.J. Hinkley, D.E. Paschon, L. Zhang, N. Psatha, Y.R. Bendana, C.M. O’Neil, A.H. Song, A.K. Mich, P.-Q. Liu, G. Lee, D.E. Bauer, M.C. Holmes, S.H. Orkin, T. Papayannopoulou, G. Stamatoyannopoulos, E.J. Rebar, P.D. Gregory, F.D. Urnov, J.A. Stamatoyannopoulos, Functional footprinting of regulatory DNA, Nat. Methods 12 (2015) 927–930. S. Neph, J. Vierstra, A.B. Stergachis, A.P. Reynolds, E. Haugen, B. Vernot, R.E. Thurman, S. John, R. Sandstrom, A.K. Johnson, M.T. Maurano, R. Humbert, E. Rynes, H. Wang, S. Vong, K. Lee, D. Bates, M. Diegel, V. Roach, D. Dunn, J. Neri, A. Schafer, R.S. Hansen, T. Kutyavin, E. Giste, M. Weaver, T. Canfield, P. Sabo, M. Zhang, G. Balasundaram, R. Byron, M.J. MacCoss, J.M. Akey, M.A. Bender, M. Groudine, R. Kaul, J.A. Stamatoyannopoulos, An expansive human regulatory lexicon encoded in transcription factor footprints, Nature 489 (2012) 83–90. M.R. Mansour, B.J. Abraham, L. Anders, A. Berezovskaya, A. Gutierrez, A.D. Durbin, J. Etchin, L. Lawton, S.E. Sallan, L.B. Silverman, M.L. Loh, S.P. Hunger, T. Sanda, R.A. Young, A.T. Look, An oncogenic super-enhancer formed through somatic mutation of a noncoding intergenic element, Science 12 (2014) 1373–1377. W.A. Flavahan, Y. Drier, B.B. Liau, S.M. Gillespie, A.S. Venteicher, A.O. Stemmer-Rachamimov, M.L. Suva, B.E. Bernstein, Insulator dysfunction and oncogene activation in IDH mutant gliomas, Nature 529 (2016) 110–114. V. Narendra, P.P. Rocha, D. An, R. Raviram, J.A. Skok, E.O. Mazzoni, D. Reinberg, CTCF establishes discrete functional chromatin domains at the Hox clusters during differentiation, Science 347 (2015) 1017–1021. H. Wang, H. Yang, C.S. Shivalila, M.M. Dawlaty, A.W. Cheng, F. Zhang, R. Jaenisch, One-step generation of mice carrying mutations in multiple genes by CRISPR/Cas-mediated genome engineering, Cell 153 (2013) 910–918. H. Yang, H. Wang, C.S. Shivalila, A.W. Cheng, L. Shi, R. Jaenisch, One-step generation of mice carrying reporter and conditional alleles by CRISPR/Casmediated genome engineering, Cell 154 (2013) 1370–1379. Y. Kim, S.-A. Cheong, J.G. Lee, S.-W. Lee, M.S. Lee, I.-J. Baek, Y.H. Sung, Generation of knockout mice by Cpf1-mediated gene targeting, Nat. Biotechnol. 34 (2016) 808–810. J.K. Hur, K. Kim, K.W. Been, G. Baek, S. Ye, J.W. Hur, S.-M. Ryu, Y.S. Lee, J.-S. Kim, Targeted mutagenesis in mice by electroporation of Cpf1 ribonucleoproteins, Nat. Biotechnol. 34 (2016) 807–808. B. Zetsche, M. Heidenreich, P. Mohanraju, I. Fedorova, J. Kneppers, E.M. DeGennaro, N. Winblad, S.R. Choudhury, O.O. Abudayyeh, J.S. Gootenberg, W. Y. Wu, D.A. Scott, K. Severinov, J. van der Oost, F. Zhang, Multiplex gene editing by CRISPR–Cpf1 using a single crRNA array, Nat. Biotechnol. (2016).

[53] N. Rajagopal, S. Srinivasan, K. Kooshesh, Y. Guo, M.D. Edwards, B. Banerjee, T. Syed, B.J.M. Emons, D.K. Gifford, R.I. Sherwood, High-throughput mapping of regulatory DNA, Nat. Biotechnol. 34 (2016) 167–174. [54] Y. Fu, J.D. Sander, D. Reyon, V.M. Cascio, J.K. Joung, Improving CRISPR-Cas nuclease specificity using truncated guide RNAs, Nat. Biotechnol. 32 (2014) 279–284. [55] D.E. Bauer, M.C. Canver, S.H. Orkin, Generation of genomic deletions in mammalian cell lines via CRISPR/Cas9, J. Vis. Exp. 95 (2015) 1–9. [56] P. Essletzbichler, T. Konopka, F. Santoro, D. Chen, B.V. Gapp, R. Kralovics, T.R. Brummelkamp, S.M.B. Nijman, T. Bürckstümmer, Megabase-scale deletion using CRISPR/Cas9 to generate a fully haploid human cell line, Genome Res. 24 (2014) 2059–2065. [57] R. Wettstein, M. Bodak, C. Ciaudo, Generation of a knockout mouse embryonic stem cell line using a paired CRISPR/Cas9 genome engineering tool, Methods Mol. Biol. (2015). [58] E. Aparicio-Prat, C. Arnan, I. Sala, N. Bosch, R. Guigó, R. Johnson, DECKO: Single-oligo, dual-CRISPR deletion of genomic elements including long noncoding RNAs, BMC Genomics 16 (2015) 846. [59] J.M. Wolfs, T.A. Hamilton, J.T. Lant, M. Laforet, J. Zhang, L.M. Salemi, G.B. Gloor, C. Schild-Poulter, D.R. Edgell, Biasing genome-editing events toward precise length deletions with an RNA-guided TevCas9 dual nuclease, Proc. Natl. Acad. Sci. (2016). [60] M. Tabebordbar, K. Zhu, J.K.W. Cheng, W.L. Chew, J.J. Widrick, W.X. Yan, C. Maesner, E.Y. Wu, R. Xiao, F.A. Ran, L. Cong, F. Zhang, L.H. Vandenberghe, G.M. Church, A.J. Wagers, In vivo gene editing in dystrophic mouse muscle and muscle stem cells, Science 351 (2016) 407–411. [61] C. Nelson, C. Hakim, D. Ousterout, P. Thakore, E. Moreb, R. Castellanos Rivera, S. Madhavan, X. Pan, F. Ran, W. Yan, A. Asokan, F. Zhang, D. Duan, C. Gersbach, In vivo genome editing improves muscle function in a mouse model of Duchenne muscular dystrophy, Science 351 (2016) 403–407. [62] C. Long, L. Amoasii, A.A. Mireault, J.R. Mcanally, H. Li, E. Sanchez-Ortiz, S. Bhattacharyya, J.M. Shelton, R. Bassel-Duby, E.N. Olson, Postnatal genome editing partially restores dystrophin expression in a mouse model of muscular dystrophy, Science 351 (2016) 400–403. [63] M.C. Canver, E.C. Smith, F. Sher, L. Pinello, N.E. Sanjana, O. Shalem, D.D. Chen, P.G. Schupp, D.S. Vinjamur, S.P. Garcia, S. Luc, R. Kurita, Y. Nakamura, Y. Fujiwara, T. Maeda, G.-C. Yuan, F. Zhang, S.H. Orkin, D.E. Bauer, BCL11A enhancer dissection by Cas9-mediated in situ saturating mutagenesis, Nature 527 (2015) 192–197. [64] D. Hnisz, J. Schuijers, C.Y. Lin, A.S. Weintraub, B.J. Abraham, T.I. Lee, J.E. Bradner, R.A. Young, Convergence of developmental and oncogenic signaling pathways at transcriptional super-enhancers, Mol. Cell 58 (2015) 362–370. [65] X. Zhang, P.S. Choi, J.M. Francis, M. Imielinski, H. Watanabe, A.D. Cherniack, M. Meyerson, Identification of focally amplified lineage-specific superenhancers in human epithelial cancers, Nat. Genet. 48 (2015) 1–8. [66] G. Andrey, K. Kraft, S. Geuer, A.J. Will, W.L. Chan, C. Paliou, M. Borschiwer, Deletions, inversions, duplications: engineering of structural variants using CRISPR/Cas in mice, Cell Rep. 10 (2015) 833–839. [67] A. Xiao, Z. Wang, Y. Hu, Y. Wu, Z. Luo, Z. Yang, Y. Zu, W. Li, P. Huang, X. Tong, Z. Zhu, S. Lin, B. Zhang, Chromosomal deletions and inversions mediated by TALENs and CRISPR/Cas in zebrafish, Nucleic Acids Res. 41 (2013) e141. [68] J. Li, J. Shou, Y. Guo, Y. Tang, Y. Wu, Z. Jia, Y. Zhai, Z. Chen, Q. Xu, Q. Wu, Efficient inversions and duplications of mammalian regulatory DNA elements and gene clusters by CRISPR/Cas9, J. Mol. Cell Biol. 7 (2015) 284–298. [69] D. Seruggia, A. Fernandez, M. Cantero, P. Pelczar, L. Montoliu, Functional validation of mouse tyrosinase non-coding regulatory DNA elements by CRISPR-Cas9-mediated mutagenesis, Nucleic Acids Res. 43 (2015) 4855– 4867. [70] Y. Guo, Q. Xu, D. Canzio, J. Shou, J. Li, D.U. Gorkin, I. Jung, H. Wu, Y. Zhai, Y. Tang, Y. Lu, Y. Wu, Z. Jia, W. Li, M.Q. Zhang, B. Ren, A.R. Krainer, T. Maniatis, Q. Wu, CRISPR inversion of ctcf sites alters genome topology and enhancer/ promoter function, Cell 162 (2015) 900–910. [71] L. Zhang, R. Jia, N.J. Palange, A.C. Satheka, J. Togo, Y. An, M. Humphrey, L. Ban, Y. Ji, H. Jin, X. Feng, Y. Zheng, Large genomic fragment deletions and insertions in mouse using CRISPR/Cas9, PLoS One 10 (2015) e0120396. [72] L. Wang, Y. Shao, Y. Guan, L. Li, L. Wu, F. Chen, M. Liu, H. Chen, Y. Ma, X. Ma, M. Liu, D. Li, Large genomic fragment deletion and functional gene cassette knock-in via Cas9 protein mediated genome editing in one-cell rodent embryos, Sci. Rep. 5 (2015) 17517. [73] C. Brandl, O. Ortiz, B. Röttig, B. Wefers, W. Wurst, R. Kühn, Creation of targeted genomic deletions using TALEN or CRISPR / Cas nuclease pairs in one-cell mouse embryos, FEBS Open Bio. 5 (2015) 26–35. [74] J. Zhou, J. Wang, B. Shen, L. Chen, Y. Su, J. Yang, W. Zhang, X. Tian, X. Huang, Dual sgRNAs facilitate CRISPR/Cas9-mediated mouse genome targeting, FEBS J. 281 (2014) 1717–1725. [75] Y. Li, A.I. Park, H. Mou, C. Colpan, A. Bizhanova, E. Akama-Garren, N. Joshi, E.A. Hendrickson, D. Feldser, H. Yin, D.G. Anderson, T. Jacks, Z. Weng, W. Xue, A versatile reporter system for CRISPR-mediated chromosomal rearrangements, Genome Biol. 16 (2015) 111. [76] X. Chen, F. Xu, C. Zhu, J. Ji, X. Zhou, X. Feng, S. Guang, Dual sgRNA-directed gene knockout using CRISPR/Cas9 technology in Caenorhabditis elegans, Sci. Rep. 4 (2014) 7581. [77] E.C. Smith, S. Luc, D.M. Croney, M.B. Woodworth, L.C. Greig, Y. Fujiwara, M. Nguyen, F. Sher, J.D. Macklis, D.E. Bauer, S.H. Orkin, Strict in vivo specificity of the Bcl11a erythroid enhancer, Blood 128 (2016) 1–4.

M.C. Canver et al. / Methods 121–122 (2017) 118–129 [78] D.G. Lupiáñez, K. Kraft, V. Heinrich, P. Krawitz, F. Brancati, E. Klopocki, D. Horn, H. Kayserili, J.M. Opitz, R. Laxova, F. Santos-Simarro, B. GilbertDussardier, L. Wittler, M. Borschiwer, S.A. Haas, M. Osterwalder, M. Franke, B. Timmermann, J. Hecht, M. Spielmann, A. Visel, S. Mundlos, Disruptions of topological chromatin domains cause pathogenic rewiring of gene-enhancer interactions, Cell 161 (2015) 1012–1025. [79] R.B. Blasco, E. Karaca, C. Ambrogio, T.-C. Cheong, E. Karayol, V.G. Minero, C. Voena, R. Chiarle, Simple and rapid in vivo generation of chromosomal rearrangements using CRISPR/Cas9 technology, Cell Rep. 9 (2014) 1219– 1227. [80] P.S. Choi, M. Meyerson, Targeted genomic rearrangements using CRISPR/Cas technology, Nat. Commun. 5 (2014) 3728. [81] D. Maddalo, E. Manchado, C.P. Concepcion, C. Bonetti, J.A. Vidigal, Y.-C. Han, P. Ogrodowski, A. Crippa, N. Rekhtman, E. de Stanchina, S.W. Lowe, A. Ventura, In vivo engineering of oncogenic chromosomal rearrangements with the CRISPR/Cas9 system, Nature 516 (2014) 423–427. [82] S. Lekomtsev, S. Aligianni, A. Lapao, T. Bürckstümmer, Efficient generation and reversion of chromosomal translocations using CRISPR/Cas technology, BMC Genomics 17 (2016) 739. [83] Y. Drier, M.J. Cotton, K.E. Williamson, S.M. Gillespie, R.J.H. Ryan, M.J. Kluk, C.D. Carey, S.J. Rodig, L.M. Sholl, A.H. Afrogheh, W.C. Faquin, L. Queimado, J. Qi, M.J. Wick, A.K. El-Naggar, J.E. Bradner, C.A. Moskaluk, J.C. Aster, B. Knoechel, B.E. Bernstein, An oncogenic MYB feedback loop drives alternate cell fates in adenoid cystic carcinoma, Nat. Genet. 48 (2016) 265–272. [84] P. Bandopadhayay, L.A. Ramkissoon, P. Jain, G. Bergthold, J. Wala, R. Zeid, S.E. Schumacher, L. Urbanski, R. O’Rourke, W.J. Gibson, K. Pelton, S.H. Ramkissoon, H.J. Han, Y. Zhu, N. Choudhari, A. Silva, K. Boucher, R.E. Henn, Y.J. Kang, D. Knoff, B.R. Paolella, A. Gladden-Young, P. Varlet, M. Pages, P.M. Horowitz, A. Federation, H. Malkin, A.A. Tracy, S. Seepo, M. Ducar, P. Van Hummelen, M. Santi, A.M. Buccoliero, M. Scagnet, D.C. Bowers, C. Giannini, S. Puget, C. Hawkins, U. Tabori, A. Klekner, L. Bognar, P.C. Burger, C. Eberhart, F.J. Rodriguez, D.A. Hill, S. Mueller, D.A. Haas-Kogan, J.J. Phillips, S. Santagata, C.D. Stiles, J.E. Bradner, N. Jabado, A. Goren, J. Grill, A.H. Ligon, L. Goumnerova, A.J. Waanders, P.B. Storm, M.W. Kieran, K.L. Ligon, R. Beroukhim, A.C. Resnick, MYB-QKI rearrangements in angiocentric glioma drive tumorigenicity through a tripartite mechanism, Nat. Genet. 48 (2016) 273–282. [85] J.A. Vidigal, A. Ventura, Rapid and efficient one-step generation of paired gRNA CRISPR-Cas9 libraries, Nat. Commun. 6 (2015) 8083. [86] A.S.L. Wong, G.C.G. Choi, C.H. Cui, G. Pregernig, P. Milani, M. Adam, S.D. Perli, S.W. Kazer, A. Gaillard, M. Hermann, A.K. Shalek, E. Fraenkel, T.K. Lu, Multiplexed barcoded CRISPR-Cas9 screening enabled by CombiGEM, Proc. Natl. Acad. Sci. U. S. A. 113 (2016) 2544–2549. [87] S. Zhu, W. Li, J. Liu, C.-H. Chen, Q. Liao, P. Xu, H. Xu, T. Xiao, Z. Cao, J. Peng, P. Yuan, M. Brown, X. Shirley Liu, W. Wei, Genome-scale deletion screening of human long non-coding RNAs using a paired-guide RNA CRISPR–Cas9 library, Nat. Biotechnol. 34 (2016) 1279–1286. [88] L.A. Miles, R.J. Garippa, J.T. Poirier, Design, execution, and analysis of pooled in vitro CRISPR/Cas9 screens, FEBS J. 283 (2016) 3170–3180. [89] J. Peng, Y. Zhou, S. Zhu, W. Wei, High-throughput screens in mammalian cells using the CRISPR-Cas9 system, FEBS J. 282 (2015) 2089–2096. [90] N.E. Sanjana, Genome-scale CRISPR pooled screens, Anal. Biochem. S0003– 2697 (2016). 30089–6.. [91] O. Shalem, N.E. Sanjana, E. Hartenian, X. Shi, D.A. Scott, T.S. Mikkelsen, D. Heckl, B.L. Ebert, D.E. Root, J.G. Doench, F. Zhang, Genome-scale CRISPR-Cas9 knockout screening in human cells, Science 343 (2014) 84–87. [92] N.E. Sanjana, O. Shalem, F. Zhang, Improved vectors and genome-wide libraries for CRISPR screening, Nat. Methods 11 (2014) 783–784. [93] Y. Zhou, S. Zhu, C. Cai, P. Yuan, C. Li, Y. Huang, W. Wei, High-throughput screening of a CRISPR/Cas9 library for functional genomics in human cells, Nature 509 (2014) 487–491. [94] J.G. Doench, N. Fusi, M. Sullender, M. Hegde, E.W. Vaimberg, K.F. Donovan, I. Smith, Z. Tothova, C. Wilen, R. Orchard, H.W. Virgin, J. Listgarten, D.E. Root, Optimized sgRNA design to maximize activity and minimize off-target effects of CRISPR-Cas9, Nat. Biotechnol. 34 (2016) 184–191. [95] S. Chen, N.E. Sanjana, K. Zheng, O. Shalem, K. Lee, X. Shi, D.A. Scott, J. Song, J.Q. Pan, R. Weissleder, H. Lee, F. Zhang, P.A. Sharp, Genome-wide CRISPR screen in a mouse model of tumor growth and metastasis, Cell 160 (2015) 1246– 1260. [96] O. Parnas, M. Jovanovic, T.M. Eisenhaure, R.H. Herbst, A. Dixit, C.J. Ye, D. Przybylski, R.J. Platt, I. Tirosh, N.E. Sanjana, O. Shalem, R. Satija, R. Raychowdhury, P. Mertins, S.A. Carr, F. Zhang, N. Hacohen, A. Regev, A Genome-wide crispr screen in primary immune cells to dissect regulatory networks, Cell 162 (2015) 675–686. [97] H. Koike-Yusa, Y. Li, E.-P. Tan, M.D.C. Velasco-Herrera, K. Yusa, Genome-wide recessive genetic screening in mammalian cells with a lentiviral CRISPRguide RNA library, Nat. Biotechnol. (2013). [98] T. Wang, J.J. Wei, D.M. Sabatini, E.S. Lander, Genetic screens in human cells using the CRISPR-Cas9 system, Science 343 (2014) 80–84. [99] O. Shalem, N.E. Sanjana, F. Zhang, High-throughput functional genomics using CRISPR–Cas9, Nat. Rev. Genet. (2015). [100] R. DeJesus, F. Moretti, G. McAllister, Z. Wang, P. Bergman, S. Liu, E. Frias, J. Alford, J.S. Reece-Hoyes, A. Lindeman, J. Kelliher, C. Russ, J. Knehr, W. Carbone, M. Beibel, G. Roma, A. Ng, J.A. Tallarico, J.A. Porter, R.J. Xavier, C. Mickanin, L. O. Murphy, G.R. Hoffman, B. Nyfeler, Functional CRISPR screening identifies the ufmylation pathway as a regulator of SQSTM1/p62, Elife 5 (2016) 1–16.

127

[101] R. Zhang, J.J. Miner, M.J. Gorman, K. Rausch, H. Ramage, J.P. White, A. Zuiani, P. Zhang, E. Fernandez, Q. Zhang, K.A. Dowd, T.C. Pierson, S. Cherry, M.S. Diamond, A CRISPR screen defines a signal peptide processing pathway required by flaviviruses, Nature 535 (2016) 164–168. [102] C.D. Marceau, A.S. Puschnik, K. Majzoub, Y.S. Ooi, S.M. Brewer, G. Fuchs, K. Swaminathan, M.A. Mata, J.E. Elias, P. Sarnow, J.E. Carette, Genetic dissection of Flaviviridae host factors through genome-scale CRISPR screens, Nature 535 (2016) 159–163. [103] J. Wallace, R. Hu, T.L. Mosbruger, T.J. Dahlem, W.Z. Stephens, D.S. Rao, J.L. Round, R.M. O’Connell, Genome-Wide CRISPR-Cas9 Screen Identifies MicroRNAs That Regulate Myeloid Leukemia Cell Growth, PLoS One 11 (2016) e0153689. [104] S. Ruiz, C. Mayor-Ruiz, V. Lafarga, M. Murga, M. Vega-Sendino, S. Ortega, O. Fernandez-Capetillo, A Genome-wide CRISPR Screen Identifies CDC25A as a Determinant of Sensitivity to ATR Inhibitors, Mol. Cell 62 (2016) 307–313. [105] T. Wang, K. Birsoy, N.W. Hughes, K.M. Krupczak, Y. Post, J.J. Wei, E.S. Lander, D.M. Sabatini, Identification and characterization of essential genes in the human genome, Science 350 (2015) 1096–1101. [106] J. Shi, E. Wang, J.P. Milazzo, Z. Wang, J.B. Kinney, C.R. Vakoc, Discovery of cancer drug targets by CRISPR-Cas9 screening of protein domains, Nat. Biotechnol. 33 (2015) 661–667. [107] J.L. Schmid-Burgk, D. Chauhan, T. Schmidt, T.S. Ebert, J. Reinhardt, E. Endl, V. Hornung, A genome-wide CRISPR (clustered regularly interspaced short palindromic repeats) screen identifies NEK7 as an essential component of NLRP3 inflammasome activation, J. Biol. Chem. 291 (2016) 103–109. [108] M.K. Kiessling, S. Schuierer, S. Stertz, M. Beibel, S. Bergling, J. Knehr, W. Carbone, C. de Vallière, J. Tchinda, T. Bouwmeester, K. Seuwen, G. Rogler, G. Roma, Identification of oncogenic driver mutations by genome-wide CRISPRCas9 dropout screening, BMC Genomics 17 (2016) 723. [109] T. Hart, M. Chandrashekhar, M. Aregger, Z. Steinhart, K.R. Brown, G. MacLeod, M. Mis, M. Zimmermann, A. Fradet-Turcotte, S. Sun, P. Mero, P. Dirks, S. Sidhu, F.P. Roth, O.S. Rissland, D. Durocher, S. Angers, J. Moffat, High-resolution CRISPR screens reveal fitness genes and genotype-specific cancer liabilities, Cell 163 (2015) 1515–1526. [110] J.D. Arroyo, A.A. Jourdain, S.E. Calvo, C.A. Ballarano, J.G. Doench, D.E. Root, V.K. Mootha, A Genome-wide CRISPR death screen identifies genes essential for oxidative phosphorylation, Cell Metab. S1550–4131 (2016) 30433–30438. [111] K. Tzelepis, H. Koike-Yusa, E. De Braekeleer, Y. Li, E. Metzakopian, O.M. Dovey, A. Mupo, V. Grinkevich, M. Li, M. Mazan, M. Gozdecka, S. Ohnishi, J. Cooper, M. Patel, T. McKerrell, B. Chen, A.F. Domingues, P. Gallipoli, S. Teichmann, H. Ponstingl, U. McDermott, J. Saez-Rodriguez, B.J.P. Huntly, F. Iorio, C. Pina, G.S. Vassiliou, K. Yusa, A CRISPR dropout screen identifies genetic vulnerabilities and therapeutic targets in acute myeloid leukemia, Cell Rep. 17 (2016) 1193– 1205. [112] Z. Steinhart, Z. Pavlovic, M. Chandrashekhar, T. Hart, X. Wang, X. Zhang, M. Robitaille, K.R. Brown, S. Jaksani, R. Overmeer, S.F. Boj, J. Adams, J. Pan, H. Clevers, S. Sidhu, J. Moffat, S. Angers, Genome-wide CRISPR screens reveal a Wnt–FZD5 signaling circuit as a druggable vulnerability of RNF43-mutant pancreatic tumors, Nat. Med. (2016). [113] R.J. Park, T. Wang, D. Koundakjian, J.F. Hultquist, P. Lamothe-Molina, B. Monel, K. Schumann, H. Yu, K.M. Krupzcak, W. Garcia-Beltran, A. PiechockaTrocha, N.J. Krogan, A. Marson, D.M. Sabatini, E.S. Lander, N. Hacohen, B.D. Walker, A genome-wide CRISPR screen identifies a restricted set of HIV host dependency factors, Nat. Genet. 49 (2016) 193–203. [114] D.R. Sen, J. Kaminski, R.A. Barnitz, M. Kurachi, U. Gerdemann, K.B. Yates, H.W. Tsao, J. Godec, M.W. LaFleur, F.D. Brown, P. Tonnerre, R.T. Chung, D.C. Tully, T.M. Allen, N. Frahm, G.M. Lauer, E.J. Wherry, N. Yosef, W.N. Haining, The epigenetic landscape of T cell exhaustion, Science (2016). [115] N. Sanjana, J. Wright, K. Zheng, O. Shalem, P. Fontanillas, J. Joung, C. Cheng, A. Regev, F. Zhang, High-resolution interrogation of functional elements in the noncoding genome, Science 353 (2016) 1545–1549. [116] Y. Diao, B. Li, Z. Meng, I. Jung, A.Y. Lee, J. Dixon, A new class of temporarily phenotypic enhancers identified by CRISPR/Cas9 mediated genetic screening, Genome Res. 26 (2016) 397–405. [117] J.B. Wright, N.E. Sanjana, CRISPR screens to discover functional noncoding elements, Trends Genet. 32 (2016) 526–529. [118] G. Korkmaz, R. Lopes, A.P. Ugalde, E. Nevedomskaya, R. Han, K. Myacheva, W. Zwart, R. Elkon, R. Agami, Functional genetic screens for enhancer elements in the human genome using CRISPR-Cas9, Nat. Biotechnol. 34 (2016) 192– 198. [119] M.C. Canver, S. Lessard, L. Pinello, Y. Wu, Y. Ilboudo, E.N. Stern, A. Needleman, F. Galactéros, C. Brugnara, A. Kutlar, D.D. Chen, P.P. Das, M.A. Cole, J. Zeng, R. Kurita, Y. Nakamura, G. Yuan, G. Lettre, D.E. Bauer, S.H. Orkin, Variant-aware saturating mutagenesis using multiple Cas9 nucleases identifies regulatory elements at trait-associated loci, Nat. Genet. (2017), http://dx.doi.org/ 10.1038/ng.3793. [120] H.K. Kim, M. Song, J. Lee, A.V. Menon, S. Jung, Y.-M. Kang, J.W. Choi, E. Woo, H. C. Koh, J.-W. Nam, H. Kim, In vivo high-throughput profiling of CRISPR–Cpf1 activity, Nat. Methods (2016). [121] A.J. Aguirre, R.M. Meyers, B.A. Weir, F. Vazquez, C.-Z. Zhang, U. Ben-David, A. Cook, G. Ha, W.F. Harrington, M.B. Doshi, M. Kost-Alimova, S. Gill, H. Xu, L.D. Ali, G. Jiang, S. Pantel, Y. Lee, A. Goodale, A.D. Cherniack, C. Oh, G. Kryukov, G. S. Cowley, L.A. Garraway, K. Stegmaier, C.W. Roberts, T.R. Golub, M. Meyerson, D.E. Root, A. Tsherniak, W.C. Hahn, Genomic copy number dictates a gene-independent cell response to CRISPR-Cas9 targeting, Cancer Discov. 2641 (2016) 617–632.

128

M.C. Canver et al. / Methods 121–122 (2017) 118–129

[122] D.M. Munoz, P.J. Cassiani, L. Li, E. Billy, J.M. Korn, M.D. Jones, J. Golji, D.A. Ruddy, K. Yu, G. McAllister, A. DeWeck, D. Abramowski, J. Wan, M.D. Shirley, S.Y. Neshat, D. Rakiec, R. de Beaumont, O. Weber, A. Kauffmann, E.R. McDonald, N. Keen, F. Hofmann, W.R. Sellers, T. Schmelzle, F. Stegmeier, M.R. Schlabach, CRISPR screens provide a comprehensive assessment of cancer vulnerabilities but generate false-positive hits for highly amplified genomic regions, Cancer Discov. 6 (2016) 900–913. [123] B.C.S. Cross, S. Lawo, C.R. Archer, J.R. Hunt, J.L. Yarker, A. Riccombeni, A.S. Little, N.J. McCarthy, J.D. Moore, Increasing the performance of pooled CRISPR–Cas9 drop-out screening, Sci. Rep. 6 (2016) 31782. [124] P.I. Thakore, A.M. D’Ippolito, L. Song, A. Safi, N.K. Shivakumar, A.M. Kabadi, T. E. Reddy, G.E. Crawford, C.A. Gersbach, A.M. D’Ippolito, L. Song, A. Safi, N.K. Shivakumar, A.M. Kabadi, T.E. Reddy, G.E. Crawford, C.A. Gersbach, Highly specific epigenome editing by CRISPR-Cas9 repressors for silencing of distal regulatory elements, Nat. Methods 12 (2015) 1143–1149. [125] C.P. Fulco, M. Munschauer, R. Anyoha, G. Munson, R. Sharon, E.M. Perez, M. Kane, B. Cleary, E.S. Lander, M. Jesse, Systematic mapping of functional enhancer-promoter connections with CRISPR interference, Science 354 (2016) 769–773. [126] L.A. Gilbert, M.A. Horlbeck, B. Adamson, J.E. Villalta, Y. Chen, E.H. Whitehead, C. Guimaraes, B. Panning, H.L. Ploegh, M.C. Bassik, L.S. Qi, M. Kampmann, J.S. Weissman, Genome-scale CRISPR-mediated control of gene repression and activation, Cell 159 (2014) 647–661. [127] S.J. Liu, S.J. Liu, M.A. Horlbeck, S.W. Cho, H.S. Birk, M. Malatesta, F.J. Attenello, J.E. Villalta, M.Y. Cho, Y. Chen, A. Mandegar, M.P. Olvera, L.A. Gilbert, B.R. Conklin, H.Y. Chang, J.S. Weissman, D.A. Lim, CRISPRi-based genome-scale identification of functional long noncoding RNA loci in human cells, Science (2016). [128] A. Amabile, A. Migliara, P. Capasso, M. Biffi, D. Cittaro, L. Naldini, A. Lombardo, Inheritable silencing of endogenous genes by hit-and-run targeted epigenetic editing, Cell 167 (2016). 219–232.e14. [129] M.L. Maeder, S.J. Linder, V.M. Cascio, Y. Fu, Q.H. Ho, J.K. Joung, CRISPR RNAguided activation of endogenous human genes, Nat. Methods 10 (2013) 977– 979. [130] P. Mali, J. Aach, P.B. Stranges, K.M. Esvelt, M. Moosburner, S. Kosuri, L. Yang, G. M. Church, CAS9 transcriptional activators for target specificity screening and paired nickases for cooperative genome engineering, Nat. Biotechnol. 31 (2013) 833–838. [131] O. Smithies, R.G. Gregg, S.S. Boggs, M.A. Koralewski, R.S. Kucherlapati, Insertion of DNA sequences into the human chromosomal beta-globin locus by homologous recombination, Nature 317 (1985) 230–234. [132] K.R. Thomas, K.R. Folger, M.R. Capecchi, High frequency targeting of genes to specific sites in the mammalian genome, Cell 44 (1986) 419–428. [133] S.L. Mansour, K.R. Thomas, M.R. Capecchi, Disruption of the proto-oncogene int-2 in mouse embryo-derived stem cells: a general strategy for targeting mutations to non-selectable genes, Nature 336 (1988) 348–352. [134] M. Bibikova, D. Carroll, D.J. Segal, J.K. Trautman, J. Smith, Y.G. Kim, S. Chandrasegaran, Stimulation of homologous recombination through targeted cleavage by chimeric nucleases, Mol. Cell. Biol. 21 (2001) 289–297. [135] F. Smih, P. Rouet, P.J. Romanienko, M. Jasin, Double-strand breaks at the target locus stimulate gene targeting in embryonic stem cells, Nucleic Acids Res. 23 (1995) 5012–5019. [136] T. Gutschner, M. Haemmerle, G. Genovese, G.F. Draetta, L. Chin, Posttranslational Regulation of Cas9 during G1 Enhances Homology-Directed Repair, Cell Rep. 14 (2016) 1555–1566. [137] T. Maruyama, S.K. Dougan, M.C. Truttmann, A.M. Bilate, J.R. Ingram, H.L. Ploegh, Increasing the efficiency of precise genome editing with CRISPR-Cas9 by inhibition of nonhomologous end joining, Nat. Biotechnol. 33 (2015) 538– 542. [138] A. Orthwein, S.M. Noordermeer, M.D. Wilson, S. Landry, R.I. Enchev, A. Sherker, M. Munro, J. Pinder, J. Salsman, G. Dellaire, B. Xia, M. Peter, D. Durocher, A mechanism for the suppression of homologous recombination in G1 cells, Nature 528 (2015) 422–426. [139] S. Lin, B. Staahl, R.K. Alla, J.A. Doudna, Enhanced homology-directed human genome engineering by controlled timing of CRISPR/Cas9 delivery, Elife 3 (2014) 1–13. [140] V.T. Chu, T. Weber, B. Wefers, W. Wurst, S. Sander, K. Rajewsky, R. Kühn, Increasing the efficiency of homology-directed repair for CRISPR-Cas9induced precise gene editing in mammalian cells, Nat. Biotechnol. 33 (2015) 543–548. [141] C. Yu, Y. Liu, T. Ma, K. Liu, S. Xu, Y. Zhang, H. Liu, M. La Russa, M. Xie, T. Ma, S. Ding, L.S. Qi, Small molecules enhance CRISPR genome editing in pluripotent stem cells, Cell Stem Cell 16 (2015) 142–147. [142] X. Ge, H. Xi, F. Yang, X. Zhi, Y. Fu, D. Chen, R.-H. Xu, G. Lin, J. Qu, J. Zhao, F. Gu, CRISPR/Cas9-AAV mediated knock-in at NRL locus in human embryonic stem cells, Mol. Ther. Acids. 5 (2016) e393. [143] J. Wang, C.M. Exline, J.J. DeClercq, G.N. Llewellyn, S.B. Hayward, P.W.-L. Li, D. A. Shivak, R.T. Surosky, P.D. Gregory, M.C. Holmes, P.M. Cannon, Homologydriven genome editing in hematopoietic stem and progenitor cells using ZFN mRNA and AAV6 donors, Nat. Biotechnol. 33 (2015) 1256–1263. [144] C.D. Richardson, G.J. Ray, M.A. DeWitt, G.L. Curie, J.E. Corn, Enhancing homology-directed genome editing by catalytically active and inactive CRISPR-Cas9 using asymmetric donor DNA, Nat. Biotechnol. (2016). [145] J. Pinder, J. Salsman, G. Dellaire, Nuclear domain ‘‘knock-in” screen for the evaluation and identification of small molecule enhancers of CRISPR-based genome editing, Nucleic Acids Res. 43 (2015) 9379–9392.

[146] D.P. Dever, R.O. Bak, A. Reinisch, J. Camarena, G. Washington, C.E. Nicolas, M. Pavel-Dinu, N. Saxena, A.B. Wilkens, S. Mantri, N. Uchida, A. Hendel, A. Narla, R. Majeti, K.I. Weinberg, M.H. Porteus, CRISPR/Cas9 b-globin gene targeting in human haematopoietic stem cells, Nature (2016). [147] D. Paquet, D. Kwart, A. Chen, A. Sproul, S. Jacob, S. Teo, K.M. Olsen, A. Gregg, S. Noggle, M. Tessier-Lavigne, Efficient introduction of specific homozygous and heterozygous mutations using CRISPR/Cas9, Nature 533 (2016) 125– 129. [148] S.E. Howden, B. Mccoll, A. Glaser, J. Vadolas, S. Petrou, M.H. Little, A.G. Elefanty, E.G. Stanley, A Cas9 variant for efficient generation of indel-free knockin or gene-corrected human pluripotent stem cells, Stem Cell Rep. 7 (2016) 508–517. [149] X. He, C. Tan, F. Wang, Y. Wang, R. Zhou, D. Cui, W. You, H. Zhao, J. Ren, B. Feng, Knock-in of large reporter genes in human cells via CRISPR/Cas9induced homology-dependent and independent DNA repair, Nucleic Acids Res. 44 (2016). [150] J.B. Renaud, C. Boix, M. Charpentier, A. De Cian, J. Cochennec, E. DuvernoisBerthet, L. Perrouault, L. Tesson, J. Edouard, R. Thinard, Y. Cherifi, S. Menoret, S. Fontanière, N. de Crozé, A. Fraichard, F. Sohm, I. Anegon, J.P. Concordet, C. Giovannangeli, Improved genome editing efficiency and flexibility using modified oligonucleotides with TALEN and CRISPR-Cas9 nucleases, Cell Rep. 14 (2016) 2263–2272. [151] T. Mikuni, J. Nishiyama, Y. Sun, N. Kamasawa, R. Yasuda, High-throughput, high-resolution mapping of protein localization in mammalian brain by in vivo genome editing, Cell 165 (2016) 1803–1817. [152] K. Nakajima, A. Kazuno, J. Kelsoe, M. Nakanishi, T. Takumi, T. Kato, Exome sequencing in the knockin mice generated using the CRISPR/Cas system, Sci. Rep. 6 (2016) 34703. [153] B.A. Parikh, D.L. Beckman, S.J. Patel, J.M. White, W.M. Yokoyama, Detailed phenotypic and molecular analyses of genetically modified mice generated by CRISPR-Cas9-mediated editing, PLoS One 10 (2015) 1–28. [154] R.J. Platt, S. Chen, Y. Zhou, M.J. Yim, L. Swiech, H.R. Kempton, J.E. Dahlman, O. Parnas, T.M. Eisenhaure, M. Jovanovic, D.B. Graham, S. Jhunjhunwala, M. Heidenreich, R.J. Xavier, R. Langer, D.G. Anderson, N. Hacohen, A. Regev, G. Feng, P.A. Sharp, F. Zhang, CRISPR-Cas9 knockin mice for genome editing and cancer modeling, Cell 159 (2014) 440–455. [155] K. Yoshimi, Y. Kunihiro, T. Kaneko, H. Nagahora, B. Voigt, T. Mashimo, SsODNmediated knock-in with CRISPR-Cas for large genomic regions in zygotes, Nat. Commun. 7 (2016) 10431. [156] P. Singh, J.C. Schimenti, The genetics of human infertility by functional interrogation of SNPs in mice, Proc. Natl. Acad. Sci. U. S. A. 112 (2015) 10431– 10436. [157] A. Raghavan, X. Wang, P. Rogov, L. Wang, X. Zhang, T.S. Mikkelsen, K. Musunuru, High-throughput screening and CRISPR-Cas9 modeling of causal lipid-associated expression quantitative trait locus variants, bioRxiv. (2016). [158] E. Crane, Q. Bian, R.P. McCord, B.R. Lajoie, B.S. Wheeler, E.J. Ralston, S. Uzawa, J. Dekker, B.J. Meyer, Condensin-driven remodelling of X chromosome topology during dosage compensation, Nature 523 (2015) 240–244. [159] S.L. Edwards, J. Beesley, J.D. French, M. Dunning, Beyond GWASs: Illuminating the dark road from association to function, Am. J. Hum. Genet. 93 (2013) 779–797. [160] S. Spisak, K. Lawrenson, Y. Fu, I. Csabai, R.T. Cottman, J.H. Seo, C. Haiman, Y. Han, R. Lenci, Q. Li, V. Tisza, Z. Szallasi, Z.T. Herbert, M. Chabot, M. Pomerantz, N. Solymosi, G.-O.E. Consortium, S.A. Gayther, J.K. Joung, M.L. Freedman, CAUSEL: an epigenome- and genome-editing pipeline for establishing function of noncoding GWAS variants, Nat. Med. 21 (2015) 1357–1363. [161] A.C. Komor, Y.B. Kim, M.S. Packer, J.A. Zuris, D.R. Liu, Programmable editing of a target base in genomic DNA without double-stranded DNA cleavage, Nature 533 (2016) 420–424. [162] K. Nishida, T. Arazoe, N. Yachie, S. Banno, M. Kakimoto, M. Tabata, M. Mochizuki, A. Miyabe, M. Araki, K.Y. Hara, Z. Shimatani, A. Kondo, Targeted nucleotide editing using hybrid prokaryotic and vertebrate adaptive immune systems, Science 102 (2016) 553–563. [163] Y. Ma, J. Zhang, W. Yin, Z. Zhang, Y. Song, X. Chang, Targeted AID-mediated mutagenesis (TAM) enables efficient genomic diversification in mammalian cells, Nat. Methods 13 (2016) 1029–1035. [164] G.T. Hess, L. Frésard, K. Han, C.H. Lee, A. Li, A. Karlene, S.B. Montgomery, M.C. Bassik, Directed evolution using dCas9-targeted somatic hypermutation in mammalian cells, Nat. Methods 13 (2016) 1036–1042. [165] G.M. Findlay, E.A. Boyle, R.J. Hause, J.C. Klein, J. Shendure, Saturation editing of genomic regions by multiplex homology-directed repair, Nature 513 (2014) 120–123. [166] T. Hashimoto, R.I. Sherwood, D.D. Kang, N. Rajagopal, A.A. Barkal, H. Zeng, B.J. M. Emons, S. Srinivasan, T. Jaakkola, D.K. Gifford, A synergistic DNA logic predicts genome-wide chromatin accessibility, Genome Res. 26 (2016) 1430– 1440. [167] S. Nakade, T. Tsubota, Y. Sakane, S. Kume, N. Sakamoto, M. Obara, T. Daimon, H. Sezutsu, T. Yamamoto, T. Sakuma, K.T. Suzuki, Microhomology-mediated end-joining-dependent integration of donor DNA in cells and animals using TALENs and CRISPR/Cas9, Nat. Commun. 5 (2014) 5560. [168] M. Maresca, V.G. Lin, N. Guo, Y. Yang, Obligate Ligation-Gated Recombination (ObLiGaRe): Custom designed nucleases mediated targeted integration through non-homologous end joining, Genome Res. 23 (2013) 539–546. [169] T.O. Auer, K. Duroure, A. De Cian, J. Concordet, F. Del Bene, Highly efficient CRISPR / Cas9-mediated knock-in in zebrafish by homology-independent DNA repair, Genome Res. 24 (2014) 142–153.

M.C. Canver et al. / Methods 121–122 (2017) 118–129 [170] K. Hoshijima, M.J. Jurynec, D.J. Grunwald, Precise editing of the zebrafish genome made simple and efficient, Dev. Cell 36 (2016) 654–667. [171] K. Suzuki, Y. Tsunekawa, R. Hernandez-Benitez, J. Wu, J. Zhu, E.J. Kim, F. Hatanaka, M. Yamamoto, T. Araoka, Z. Li, M. Kurita, T. Hishida, M. Li, E. Aizawa, S. Guo, S. Chen, A. Goebl, R.D. Soligalla, J. Qu, T. Jiang, X. Fu, M. Jafari, C.R. Esteban, W.T. Berggren, J. Lajara, E. Nuñez-Delicado, P. Guillen, J.M. Campistol, F. Matsuzaki, G.-H. Liu, P. Magistretti, K. Zhang, E.M. Callaway, K. Zhang, J.C.I. Belmonte, In vivo genome editing via CRISPR/Cas9 mediated homology-independent targeted integration, Nature 540 (2016) 144–149. [172] A.W. Cheng, H. Wang, H. Yang, L. Shi, Y. Katz, T.W. Theunissen, S. Rangarajan, C.S. Shivalila, D.B. Dadon, R. Jaenisch, Multiplexed activation of endogenous genes by CRISPR-on, an RNA-guided transcriptional activator system, Cell Res. 23 (2013) 1163–1171. [173] S. Konermann, M.D. Brigham, A.E. Trevino, J. Joung, O.O. Abudayyeh, C. Barcena, P.D. Hsu, N. Habib, J.S. Gootenberg, H. Nishimasu, O. Nureki, Genome-scale transcriptional activation by an engineered CRISPR-Cas9 complex, Nature 517 (2015) 583–588. [174] A. Chavez, M. Tuttle, B.W. Pruitt, E.-C. Ben, R. Chari, T.-O. Dmitry, S.J. Haque, R. J. Cecchi, E.J.K. Kowal, J. Buchthal, B.E. Housden, N. Perrimon, J.J. Collins, G. Church, B. Ewen-Campen, R. Chari, D. Ter-Ovanesyan, S.J. Haque, R.J. Cecchi, E.J.K. Kowal, J. Buchthal, B.E. Housden, N. Perrimon, J.J. Collins, G. Church, Comparison of Cas9 activators in multiple species, Nat. Methods 13 (2016) 563–567. [175] D.R. Simeonov, B.G. Gowen, M. Boontanrart, T. Roth, Y. Lee, A. Chan, M.L. Nguyen, R.E. Gate, M. Subramaniam, J.M. Woo, T. Mitros, G.J. Ray, N.L. Bray, G. L. Curie, N. Naddaf, E. Boyer, F. Van Gool, K. Schumann, M.J. Daly, K.K. Fahr, C. Ye, J.A. Bluestone, M.S. Anderson, J.E. Corn, A. Marson, Discovery of an autoimmunity-associated IL2RA enhancer by unbiased targeting of transcriptional activation, bioRxiv. (2016). [176] I.B. Hilton, A.M. D’Ippolito, C.M. Vockley, P.I. Thakore, G.E. Crawford, T.E. Reddy, C.A. Gersbach, Epigenome editing by a CRISPR-Cas9-based acetyltransferase activates genes from promoters and enhancers, Nat. Biotechnol. 33 (2015) 510–517. [177] J.G. Doench, E. Hartenian, D.B. Graham, Z. Tothova, M. Hegde, I. Smith, M. Sullender, B.L. Ebert, R.J. Xavier, D.E. Root, Rational design of highly active sgRNAs for CRISPR-Cas9-mediated gene inactivation, Nat. Biotechnol. 32 (2014) 1262–1267. [178] M. Haeussler, K. Schönig, H. Eckert, A. Eschstruth, J. Mianné, J.-B. Renaud, S. Schneider-Maunoury, A. Shkumatava, L. Teboul, J. Kent, J.-S. Joly, J.-P. Concordet, Evaluation of off-target and on-target scoring algorithms and integration into the guide RNA selection tool CRISPOR, Genome Biol. 17 (2016) 148. [179] H. Xu, T. Xiao, C.-H. Chen, W. Li, C.A. Meyer, Q. Wu, D. Wu, L. Cong, F. Zhang, J. S. Liu, M. Brown, X.S. Liu, Sequence determinants of improved CRISPR sgRNA design, Genome Res. 25 (2015) 1147–1157. [180] N. Fusi, I. Smith, J. Doench, J. Listgarten, In Silico Predictive Modeling of CRISPR/Cas9 guide efficiency, bioRxiv. (2015). [181] R. Chari, P. Mali, M. Moosburner, G.M. Church, Unraveling CRISPR-Cas9 genome engineering parameters via a library-on-library approach, Nat. Methods 12 (2015) 823–826. [182] M.A. Moreno-Mateos, C.E. Vejnar, J. Beaudoin, J.P. Fernandez, E.K. Mis, M.K. Khokha, A.J. Giraldez, CRISPRscan: designing highly efficient sgRNAs for CRISPR-Cas9 targeting in vivo, Nat. Methods 12 (2015) 982–988. [183] B.E. Housden, A.J. Valvezan, C. Kelley, R. Sopko, Y. Hu, C. Roesel, S. Lin, M. Buckner, R. Tao, B. Yilmazel, S.E. Mohr, B.D. Manning, N. Perrimon, Identification of potential drug targets for tuberous sclerosis complex by synthetic screens combining CRISPR-based knockouts with RNAi, Sci. Signal. 8 (2015). rs9.. [184] S. Bae, J. Kweon, H.S. Kim, J.-S. Kim, Microhomology-based choice of Cas9 nuclease target sites, Nat. Methods 11 (2014) 705–706. [185] B. Farboud, B.J. Meyer, Dramatic enhancement of genome editing by CRISPR/cas9 through improved guide RNA design, Genetics 199 (2015) 959–971. [186] X. Ren, Z. Yang, J. Xu, J. Sun, D. Mao, Y. Hu, S.J. Yang, H.H. Qiao, X. Wang, Q. Hu, P. Deng, L.P. Liu, J.Y. Ji, J.B. Li, J.Q. Ni, Enhanced specificity and efficiency of the CRISPR/Cas9 system with optimized sgRNA parameters in Drosophila, Cell Rep. 9 (2014) 1151–1162. [187] V.T. Chu, R. Graf, T. Wirtz, T. Weber, Efficient CRISPR-mediated mutagenesis in primary immune cells using CrispRGold and a new Cas9 transgenic mouse line, (2016). [188] C.E. Vejnar, M.A. Moreno-Mateos, D. Cifuentes, A.A. Bazzini, A.J. Giraldez, Optimization strategies for the CRISPR–Cas9 genome-editing system, Cold Spring Harb. Protoc. 2016 (2016) 829–833.

129

[189] M.A. Horlbeck, L.A. Gilbert, J.E. Villalta, B. Adamson, R.A. Pak, Y. Chen, A.P. Fields, C.Y. Park, J.E. Corn, M. Kampmann, J.S. Weissman, Compact and highly active next-generation libraries for CRISPR-mediated gene repression and activation, Elife 5 (2016) e19760. [190] M. Stemmer, T. Thumberger, M. Del Sol Keyer, J. Wittbrodt, J.L. Mateo, CCTop: An intuitive, flexible and reliable CRISPR/Cas9 target prediction tool, PLoS One 10 (2015) 1–11. [191] S.Q. Tsai, Z. Zheng, N.T. Nguyen, M. Liebers, V.V. Topkar, V. Thapar, N. Wyvekens, C. Khayter, A.J. Iafrate, L.P. Le, M.J. Aryee, J.K. Joung, GUIDE-seq enables genome-wide profiling of off-target cleavage by CRISPR-Cas nucleases, Nat. Biotechnol. 33 (2015) 187–197. [192] R.L. Frock, J. Hu, R.M. Meyers, Y. Ho, E. Kii, F.W. Alt, Genome-wide detection of DNA double-stranded breaks induced by engineered nucleases, Nat. Biotechnol. 33 (2014) 179–186. [193] X.-H. Zhang, L.Y. Tee, X.-G. Wang, Q.-S. Huang, S.-H. Yang, Off-target Effects in CRISPR/Cas9-mediated genome engineering, Mol. Ther. Nucleic Acids 4 (2015). e264.. [194] J. Tycko, V.E. Myer, P.D. Hsu, Methods for optimizing CRISPR-Cas9 genome editing specificity, Mol. Cell 63 (2016) 355–370. [195] J.P. Guilinger, D.B. Thompson, D.R. Liu, Fusion of catalytically inactive Cas9 to FokI nuclease improves the specificity of genome modification, Nat. Biotechnol. (2014). [196] B.P. Kleinstiver, V. Pattanayak, M.S. Prew, S.Q. Tsai, N.T. Nguyen, Z. Zheng, J. Keith Joung, High-fidelity CRISPR–Cas9 nucleases with no detectable genome-wide off-target effects, Nature 529 (2016) 490–495. [197] I.M. Slaymaker, L. Gao, B. Zetsche, D.A. Scott, W.X. Yan, F. Zhang, Rationally engineered Cas9 nucleases with improved specificity, Science 351 (2016) 84– 88. [198] B.P. Kleinstiver, S.Q. Tsai, M.S. Prew, N.T. Nguyen, M.M. Welch, J.M. Lopez, Z.R. McCaw, M.J. Aryee, J.K. Joung, Genome-wide specificities of CRISPR-Cas Cpf1 nucleases in human cells, Nat. Biotechnol. 34 (2016) 869–874. [199] D. Kim, J. Kim, J.K. Hur, K.W. Been, S.-H. Yoon, J.-S. Kim, Genome-wide analysis reveals specificities of Cpf1 endonucleases in human cells, Nat. Biotechnol. 34 (2016) 863–868. [200] Y. Gao, X. Xiong, S. Wong, E.J. Charles, W.A. Lim, L.S. Qi, Complex transcriptional modulation with orthogonal and inducible dCas9 regulators, Nat. Methods 13 (2016) 1043–1049. [201] K.I. Liu, M.N. Bin Ramli, C.W.A. Woo, Y. Wang, T. Zhao, X. Zhang, G.R.D. Yim, B. Y. Chong, A. Gowher, M.Z.H. Chua, J. Jung, J.H.J. Lee, M.H. Tan, A chemicalinducible CRISPR–Cas9 system for rapid control of genome editing, Nat. Chem. Biol. (2016). [202] B. Maji, C.L. Moore, B. Zetsche, S.E. Volz, F. Zhang, M.D. Shoulders, A. Choudhary, Multidimensional chemical control of CRISPR-Cas9, Nat. Chem. Biol. 5–7 (2016). [203] A. Pawluk, N. Amrani, Y. Zhang, B. Garcia, Y. Hidalgo-Reyes, J. Lee, A. Edraki, M. Shah, E.J. Sontheimer, K.L. Maxwell, A.R. Davidson, Naturally Occurring Off-Switches for CRISPR-Cas9, Cell 167 (2016). 1829–1838.e9. [204] D.A. Jaitin, A. Weiner, I. Yofe, D. Lara-Astiaso, H. Keren-Shaul, E. David, T.M. Salame, A. Tanay, A. van Oudenaarden, I. Amit, Dissecting Immune Circuits by Linking CRISPR-Pooled Screens with Single-Cell RNA-Seq, Cell 167 (2016). 1883–1896.e15. [205] A. Dixit, O. Parnas, B. Li, J. Chen, C.P. Fulco, L. Jerby-Arnon, N.D. Marjanovic, D. Dionne, T. Burks, R. Raychowdhury, B. Adamson, T.M. Norman, E.S. Lander, J.S. Weissman, N. Friedman, A. Regev, Perturb-Seq: Dissecting Molecular Circuits with Scalable Single-Cell RNA Profiling of Pooled Genetic Screens, Cell 167 (2016). 1853–1866.e17. [206] B. Adamson, T.M. Norman, M. Jost, M.Y. Cho, J.K. Nuñez, Y. Chen, J.E. Villalta, L. A. Gilbert, M.A. Horlbeck, M.Y. Hein, R.A. Pak, A.N. Gray, C.A. Gross, A. Dixit, O. Parnas, A. Regev, J.S. Weissman, A Multiplexed Single-Cell CRISPR Screening Platform Enables Systematic Dissection of the Unfolded Protein Response, Cell 167 (2016). 1867–1882.e21. [207] J.G. Zalatan, M.E. Lee, R. Almeida, L.A. Gilbert, E.H. Whitehead, M. La Russa, J.C. Tsai, J.S. Weissman, J.E. Dueber, L.S. Qi, W.A. Lim, Engineering Complex Synthetic Transcriptional Programs with CRISPR RNA Scaffolds, Cell 160 (2015) 339–350. [208] D. Cox, R.J. Platt, F. Zhang, Therapeutic genome editing : prospects and challenges, Nat. Med. 21 (2015) 121–131. [209] M.C. Canver, S.H. Orkin, Customizing the genome as therapy for the bhemoglobinopathies, Blood 127 (2016) 2536–2545. [210] M.L. Maeder, C.A. Gersbach, Genome-editing Technologies for Gene and Cell Therapy, Mol. Ther. 24 (2016) 430–446.

Suggest Documents