Open access peer-reviewed chapter

The Prominent Characteristics of the Effective sgRNA for a Precise CRISPR Genome Editing

Written By

Reza Mohammadhassan, Sara Tutunchi, Negar Nasehi, Fatemeh Goudarziasl and Lena Mahya

Submitted: 25 June 2022 Reviewed: 22 July 2022 Published: 19 August 2022

DOI: 10.5772/intechopen.106711

From the Edited Volume

CRISPR Technology - Recent Advances

Edited by Yuan-Chuan Chen

Chapter metrics overview

201 Chapter Downloads

View Full Metrics

Abstract

Clustered regularly interspaced short palindromic repeats (CRISPRs) technique is the most effective and novelist technique for genome editing. CRISPR mechanism has been widely developed for gene editing, gene silencing, high-specific regulation of the transcription, and reducing off-target effects through double-strand breaks (DSBs) in the genomic DNA and then modifying nucleotide sequences of the target gene in diverse plant and animal species. However, the application may be restricted by a high rate of off-target effects. So, there are many studies on designing precise single-guide RNAs (sgRNAs) to minimize off-target effects. Thus, the high-efficiency design of a specific sgRNA is critical. First, in the chapter, the sgRNA origin and different types of gRNA will be outlined. Then, the off-target effect will be described. Next, the remarkable characteristics of the sgRNA will be highlighted to improve precise gene editing. Finally, some popular in silico tools will be introduced for designing sgRNA.

Keywords

  • sgRNA
  • guide RNA
  • crRNA
  • tracrRNA
  • off-target effect
  • designing tools
  • CRISPR/Cas

1. Introduction

Clustered regularly interspaced short palindromic repeats (CRISPRs) and their CRISPR-associated (Cas) proteins system is an effective immune system among bacteria and archaea. This system was first discovered in the E.coli genome. CRISPR/Cas is an acquired immunity mechanism in many bacteria and archaea against the genome of the infection factors such as viruses and plasmids [1]. CRISPR/Cas system is classified into three major groups (I, II, and III) with a specific functional mechanism and gene family encoding the specific Cas proteins. Types I and III apply several Cas proteins for endonuclease activity, while type II uses only one protein (Cas9) [2]. Evolutionary competition between the pathogens and host in the CRISPR/Cas system shows a very high variable rate in structures and functions. So, recent classification has stated that the CRISPR system has been categorized into two classes (I and II) and six types (I–VI) [3].

Most studies on genome engineering have been performed in system type II, derived from Streptococcus pyogenes (SpCas9). The advantage of system type II is that it needs only one protein (Cas9) for endonuclease activity. However, this system also needs the types of RNA, including CRISPR RNA (crRNA), which functions as a Cas protein guide for pairing with the target genome sequences, and trans-activating CRISPR RNA (tracrRNA), which play critical roles in crRNA maturation and directing Cas9 to the desired site [4, 5]. Merging of tracrRNA:crRNA sequences as a chimeric sequence known as single-guide RNA (sgRNA), which covers the features of both RNA types, makes it suitable and applicable for genome editing [6].

All CRISPR sites cover consecutive and spacer repetitions. These consecutive repetitions include identical sequences, while the spacer sequences originate from the genome of foreign factors [7, 8]. CRISPR sites and Cas proteins develop acquired immunity against the invading DNA. Suppose a microorganism survives the invasion of a pathogen. In that case, the integrated CRISPR system will be able to incorporate a piece of the pathogen DNA into its genome and then use it to fight against subsequent infections. This bacterial immune system degenerates the phage genome by integrating short fragments of the phage DNA in the spacer region of the CRISPR sites and transcribing the spacers (known as crRNAs) with associated Cas endonuclease in subsequent infections [9, 10, 11].

Briefly, the CRISPR system, as the RNA-mediated immune system in prokaryotes (bacteria and archaea), functions in four stages (Figure 1):

  1. Admission: short viral or plasmid DNA fragments are recognized and then inserted as a spacer between two consecutive repetitions into the CRISPR sites [12, 13].

  2. Expression: the CRISPR sites are transcribed to a long precursor crRNA (pre-crRNA) containing a complete array of CRISPR repetitions and sequences derived from infectious factors [14].

  3. Insertion: Pre-crRNA is cleaved by a special endoribonuclease into small guide sequences known as crRNA [15].

  4. Targeting: The crRNAs guide Cas endonuclease to cleave complementary DNA or RNA sequences flanked by a protospacer adjacent motif (PAM) (for DNA targets) or a protospacer flanking sequence (usually known as PFS) without significant complement to the crRNA repeats (for RNA targets) [16, 17].

Figure 1.

RNA-mediated CRISPR immune system in.

The classification represents the evolution of subtype-specific molecular defensive mechanisms for crRNA expression and maturation, as well as the inhibition of infectious factors [18]. The main known vital component of the CRISPR/Cas system is crRNA, which is common in types I and III. Pre-crRNAs are initially cleaved within the repeats by a Cas6 endoribonucleases family, and then intermediate crRNAs are further matured to generate shorter repeat-spacer crRNAs in type III. In both type I and type III, mature crRNAs direct a complex of multiple Cas proteins to the cognate-invading nucleic acids. Then, the target nucleic acids are cleaved by a Cas endonuclease of the ribonucleoprotein complex [19, 20, 21].

Pre-crRNA processing necessitates base-pairing of each pre-crRNA repeat with tracrRNA, a small noncoding RNA encoded near the Cas genes and spacer array [7]. The base-pairing drives cleavage and binding by RNase III and Cas9, respectively. Then, the crRNA:tracrRNA complex can direct Cas9 to bind target DNA sequences by matching PAM [16, 17]. In addition, type II CRISPR-Cas systems continuously utilize the crRNA:tracrRNA complex to identify and cleave double-stranded DNA (dsDNA). Recognition is driven by base-pairing between the guide sequence and the RNA in the crRNA:tracrRNA complex (Figure 2) [22, 23]. The tracrRNA is a common component of CRISPR-Cas systems among all three subtypes of type II systems and is required for crRNA biosynthesis [24]. These basic principles were discovered by following efforts to characterize tracrRNAs and crRNA biogenesis [25]. The biogenesis of crRNA, the structures of the crRNA:tracrRNA complex, and tracrRNA genomic location are variable among these subtypes [26]. The tracrRNA discovery as a key factor of crRNA biosynthesis allowed the sgRNA to be invented and Cas9 to be adopted as the core component of CRISPR technology [27].

Figure 2.

Schematic structure of the crRNA:tracrRNA complex.

The major function of the sgRNA is efficiently the target region detection through the PAM sequence to edit a gene precisely. However, two vital challenges include efficacy and specificity for designing an effective sgRNA [28]. According to the significance of the sgRNA function, the role and designing tools of sgRNA will be outlined in the following.

Advertisement

2. The role of sgRNA in CRISPR technology

CRISPR/Cas genome editing technology can cause double-stranded DNA breaks (DSBs) in predefined genomic loci [29]. These DSBs are then repaired by the DNA repair systems of the target organism, which inherently can cause a mutation in the target gene. Despite the general processes driving all these genome editing systems being similar, CRISPR/Cas technology has emerged as the preferred technique due to its easy usage, low cost, outstanding adaptability, and ability to target several genes at once [30, 31, 32].

This technology consists of a Cas endonuclease, responsible for eliciting the DSB, and a short noncoding about sgRNA (20-nt), directing Cas to the correct genomic region for targeted genome editing. A chimeric gRNA (complementary to the target area) and trans-activating CRISPR-RNA are usually included in this sgRNA. Most Cas systems require the predesigned sgRNA to anneal immediately upstream of a PAM, which in the case of SpCas9 (the most extensively employed Cas protein for genome editing) is 5` NGG3` [4, 5, 33]. In these cases, the PAM is required to cleave target DNA around 3-nt upstream of this region. DSBs are repaired by either nonhomologous end joining (NHEJ) or homology-directed repair (HDR) as two central intrinsic DNA repair systems [34]. The error-prone nature of NHEJ is the main DNA repair route in species and the most common and straightforward pathway in genome editing, causing small insertions or deletions (indels) to interrupt the target sequence (Figure 3) [35].

Figure 3.

CRISPR/Cas9 technology for gene editing. The Cas9 DNA endonuclease is recruited by a single-guide RNA (sgRNA) that detects a genomic sequence followed by a 5′-NGG-3′ PAM motif.

CRISPR/Cas technology relies on DNA-RNA interaction as well as simple design of RNA molecule for each specific sequence. However, protein-DNA interaction also depends on zinc-finger nucleases (ZFNs) and transcription activator-like effector nucleases (TALENs). So there is a required reconstruction for each target DNA sequence. This is a significant benefit of CRISPR/Cas technology [36]. In addition, the technology has three other advantages over TALENs and ZFNs, including the following (Table 1):

  1. Simplicity: The 20 nt sequence of sgRNA can be simply designed to target any sequence [5].

  2. Multifunction: As this technology’s main advantage over TALENs and ZFNs, several sgRNAs can function differently and simultaneously on variable genomic sites [37].

  3. Non-specificity to the target DNA methylation: the CRISPR/Cas technology can edit some genome sites which are highly regulated by epigenetic changes [38], especially in plants that contain around 70% of CpG/CpNpG sites and CpG islands methylated in the proximal exons promoter [39].

2.1 The role of PAM

Although specific targeting significantly depends on the sgRNA sequence, PAM-specific sequence plays a crucial role in the effective enzymatic activity of Cas endonuclease. In this system, Cas9 endonuclease can cleave any genome sequence immediately located on five nucleotides (nt) downstream of PAM sequence [25]. PAM sequences recognized by SpCas9 and StCas9 (from Streptococcus thermophilus) is, respectively, 3`-NGG-5` and 5`-NNAGAAW-3` (Table 2) [40, 41]. The SpCas9-sgRNA complex first seeks the complement sequence of PAM in the target genome and then the sgRNA base pairs to the target DNA, then the DNA is cleaved by SpCas9 to create DSB. Generally, the length of the DNA detection sequence in the crRNA region is 20 nt, though the more base pairs bind between RNA and DNA, the more the specificity of the sgRNA function can enhance. Hence, the 20 nt sequence of sgRNA and 3 nt in PAM play key roles in the specific targeting of CRISPR/Cas technology. However, there are some limitations in using the 3`-NGG-5` motif, particularly at high AT sequences of the target genome [42, 43].

SystemZFNsTALENsCRISPR
FunctionCleavage mediated by DNA-protein interactionCleavage mediated by DNA-protein interactionCleavage mediated by DNA-RNA interaction
Nuclease designing and assemblyHard and costlyMostly possible in the laboratory, but highly difficultEasy
Designing efficiencyLowHighHigh
Assembly efficiencyVariableHigh (%99<)High (%90<)
Targeting rangeRestricted, because of dependence on ZF modulesUnrestricted because of independence on PAMRestricted by PAM, but generally unrestricted
Off-target effectsYesYesYes
Sensitivity to DNA methylationUndefinedSensitive to CpG methylationNonsensitive to CpG methylation
High operating powerNoRestrictedPossible

Table 1.

Comparison between genome editing systems.

Bacterial speciesEndonucleasePAM-specific sequence
Streptococcus pyogenesSpCas9NGG
SpCas9 D1135ENAG
SpCas9 VRERNGCG
SpCas9 EQRNGAG
SpCas9 VQRNGNG/ NGAN
Staphylococcus aureusSaCas9NNGRR(N)/ NNGRRT
Neisseria meningitidisNmCas9NNNNGATT
Streptococcus thermophilusStCas9NNAGAAW
Treponema denticolaTdCas9NAAAAC
S. aureusSaCas12aTTN

Table 2.

Different PAM sequences of some Cas endonucleases and their origins.

N = A,T,G,C; R = G,A; and W = A, T.

2.2 Variety of sgRNA types

According to the development of the CRISPR system as technology and sgRNA invention, there are some improvements of sgRNA to enhance the efficiency, precision, and specificity of the genome editing technology [44]. The improvements include as follows:

  1. Truncated guide RNA (tuRNA): the RNA contains a homologous 17 nt sequence to the target gene. Hence, the specificity of Cas endonuclease activity can increase by reducing off-target sequences [45].

  2. Polycistronic tRNA-gRNA (PTG/Cas9): in the system, the RNA is a frequent repetition of the tRNA-gRNA units and target-specific spacer sequences for targeting several sites in the genome sequence [46]. After PTG transcription, the primary copy is matured as sgRNAs through tRNA processing system and RNaseP and RNaseZ activity. In addition, the sgRNAs can interact with Cas endonuclease to target multiple genes specifically and simultaneously [47].

2.3 Off-target effect

Off-target mutation is a major challenge in CRISPR/Cas technology [48]. If the gRNA sequence contains less than three heterologous nucleotides to an off-target region, off-target effects will be observed [49]. Studies have indicated that mismatched pairs at the end of the 3` terminal of the target sequence are not tolerated (typically 8–14 nucleotides upstream of PAM sequence). In contrast, the mismatched pairs at the 5` terminal of the target sequence are more tolerable [50]. The sgRNA/Cas9 also can influence the off-target effects [49].

Generally, although Cas9 protein can be differently used according to high endonuclease activity and the wide targeting range of the enzyme, the high molecular weight of the Cas9 endonuclease and off-target effects can restrict the popularity of the enzyme. Nevertheless, some variants of Cas9, such as SpCas9-HF and eSpCas9 [51, 52], can be mutated. So, the mutation can reduce the nonspecific interaction between the Cas9 protein and the target sequence. Digenome-seq, GUIDE-Seq (Genome-wide, Unbiased Identification of DSBs Enabled by Sequencing), and HTGTS (High-Throughput Genome-Wide Translocation Sequencing) can be employed for detecting off-target regions [53]. However, precisely designing sgRNA can significantly decrease the rate of the off-target effects [44].

Advertisement

3. Characteristics of an effective sgRNA

In addition to directing Cas endonuclease, sgRNA can stimulate Cas endonuclease activity. These functions of sgRNA can clarify how to tackle on-target effects [54]. It has been demonstrated that the proximal and distal ends of the PAM sequence are highly responsible for improving on-target effects. Besides, genomic frameworks of the target sequence, GC content, sgRNA length, and secondary structure play significant roles in enhancing the on-target rate [44]. Also, 5` terminal 20 nt of the sgRNA is highly efficient for on-target efficiency. However, there is insufficient information on the correlation between structure and sequence properties of sgRNA influencing the on-target effect [55]. At least, enhancement of the on-target rate can improve the efficiency of the CRISPR technology and facilitate the statistical interpretation of the edited genes rate [56].

3.1 GC content of the sgRNA

As mentioned above, the GC content of the sgRNA is closely related to the on-target rate and the efficiency of the CRISPR gene editing. It has been indicated that too high or low GC content is unsuitable for achieving high rate of the on-target effects [57]. The knockout effectiveness of the CRISPR/Cas9 system was significantly improved by changing the sgRNA structure by expanding the duplex length (about 5 nt) and replacing the fourth T by C or G [58]. Many studies have reported that GC content plays a key role in improving the knockout efficiency of the CRISPR/Cas9 system. The effective rate of GC content is 40–60 percent. It is also recommended that sgRNA containing 50 percent GC content is efficient for CRISPR gene editing [59, 60, 61]. However, some studies suggest higher than 60 percent GC content for each organism, such as Escherichia coli (62.5%) and Vitis vinifera (65%) [62, 63]. In addition to GC content, purine residues and curvature in positions C3 and C16 of sgRNA could be effective in improving on-target activity [64]. It has also been reported that sgRNAs containing 4 GC in the 6 nt close to the PAM sequence can effectively reduce off-target effects [65].

3.2 The sgRNA length

The most common sgRNA length is about 100 nt. Therefore, 5` terminal 20 nt of the sgRNA can be designed as the complement sequence of the target gene to direct precisely Cas endonuclease for achieving effective gene editing [66]. According to several studies, the less the sgRNA length is, the higher the rate of off-target effects increase [67, 68, 69]. However, as mentioned above, 17 nt length for tuRNA can be highly effective in reducing off-target effects. However, the length of the sgRNA recognition site is less than 15 nt, and Cas endonuclease will not show any activity [70, 71]. In addition to the sgRNA length, the efficiency and specificity of Cas cleavage activity in the targeting sequence are significantly influenced by the distance between the PAM site and the start codon [72].

3.3 The sgRNA secondary structure

The sgRNA secondary structure is highly responsible for effective Cas-target sequence binding [73]. There are also many reports to indicate that the presence of the quad stem-loop structure of sgRNA is a key factor in improving the efficiency of the riboprotein function. The repeat and anti-repeat region (stem-loop RAR, GAAA) can activate sgRNA processing before Cas-sgRNA binding. Besides, loops 2 (GAAA) and 3 (AGU) are demanded to create a stable riboprotein complex, but there is no report on the possible loop 1 role in sgRNA efficiency [74, 75, 76, 77]. Besides, the hairpin structure of the sgRNA, particularly the inner side of the hairpin, plays a key role in cleaving target DNA by Cas9/sgRNA. In fact, the hairpin structure can provide a suitable conformation to bind Cas9 enzyme. If the loop-stem structure is elongated, the gene editing efficiency will be enhanced [78, 79]. It has been indicated that CRISPR efficiency can be improved when an engineered hairpin structure is inserted into the spacer region of sgRNA. The modified sgRNA can positively influence Cas-mediated transactivation and improve the function of the five different Cas9 and Cas12a variants. The evidence can demonstrate the effect of sgRNA secondary structure on the success rate of gene editing [80, 81]. There is also a correlation between the sgRNA secondary structure and the efficiency of the Cas9-mediated CRISPR [55]. It has been reported that sgRNA refolding can refine the destructive bonds of the deactivated sgRNAs. Also, heating or slowly chilling can thermodynamically activate these sgRNAs to improve Cas9 cleavage activity. At least, the sgRNA secondary structure can change the guide sequence activity and deactivated sgRNA can be recovered by refolding [82, 83].

3.4 The sgRNA sequence

In addition to these criteria, the sgRNA sequence features can show different efficiency levels. It has been reported that the functional sgRNA can be significantly accessed at certain nucleotide positions more than nonfunctional sgRNA. Particularly, 3` terminal nucleotides (positions 18–20) of the sgRNA can highly make a prominent difference in accessibility [84, 85]. The sgRNA 3` terminus, known as the seed region, is a key player in recognizing the target sequence. Therefore, accessibility of the final three bases of the seed region is a remarkable characteristic in distinguishing functional sgRNAs from nonfunctional ones based on structural analysis [86, 87]. Also, G at the 5` terminus of the spacer is demanded in the non-ribosomal and ribosomal complexes. Besides, G is extremely proper at positions −1 and − 2, close to PAM associated with the sequence preference for loading Cas9 endonuclease [88]. According to the finding that multiple U in the spacer results in low sgRNA expression, T is not preferred at the four positions nearest to the PAM. The downstream nucleotides of the PAM cooperate with the efficiency of sgRNA, while upstream sequences of the spacer do not show significant effects. The early termination of sgRNA transcription is mostly responsible for the reduced sgRNA expression rates caused by the high frequency of nonconsecutive T clusters in the protospacer. Cytosine is favored at the −3 position as the cleavage site of the sgRNA/Cas9 complex.

Additionally, guanines are favored from positions −14 to −17, while adenines are favored positions −5 to −12 [89, 90, 91, 92, 93]. Most molecular characteristics that govern sgRNA stability, loading, and targeting in vivo are yet unknown. While variable Cas9 off-target binding, positioning of the nucleosome, and sgRNA loading are not key factors, adenine depletion and guanine enrichment improved sgRNA activity and stability. There is also a close correlation between sgRNA efficiency and guanine enrichment PAM-proximal site, supposedly caused by G-quadruple structure increasing sgRNA stability [94, 95, 96, 97].

Advertisement

4. Computational tools for designing sgRNA

Generally, the potential off-target effect is still a vital concern for several applications of CRISPR technology. There are many strategies including Cas endonucleases engineering, transcriptome analysis, tunable systems (small-molecule induction of Cas9, light-activated and intein-inactivated Cas9), functional screening after dCas9 treatment, direct delivery (RNP complex), truncated sgRNAs (small-guide RNAs), and separate Cas9-binding approaches (paired Cas9 nickase) to reduce the off-target activity in CRISPR/Cas gene editing (Table 3) [105, 106, 107]. However, designing sgRNA can be the most simple, effective, low-cost, and time-saving approach [28].

There are some key factors for designing an efficient sgRNA for CRISPR editing and reducing off-target effects. First, the GC contact should be 40–80 per cent, although a higher percentage is more desirable and beneficial [28]. Second, the sgRNA length needs to be 17–20 nt, depending on the used Cas enzyme. The shorter the sequence is designed, the less the off-target effects are observed; however, too short sequences can increase the off-target effects [71]. As the third factor, off-target effects may result from mismatches between sgRNA and the target sequence, according to the mismatch numbers and positions [96]. ΔG calculation provides a significant benefit to assess sgRNA-DNA binding potential. The ΔG of a highly effective gRNA is from −64.53 to −47.09 kcal/mol, but higher ΔG can increase mismatching rates, causing a high rate of the off-target effect [93]. Although the more the negative ΔG would be, the more stable the secondary structure of the sgRNA is observed, higher ΔG (~ −30 to −20 kcal/mol) can positively influence sgRNA transcription and practically make more effective RNA-RNA binding for a more functional secondary structure of the sgRNA [28].

According to all these criteria, designing sgRNA is a crucial concern in CRISPR technology [72]. As CRISPR/Cas contain two key players, including Cas endonuclease and sgRNA, to cooperate with genome editing, the development of each component could be beneficial to enhancing CRISPR/Cas editing. However, enzyme engineering is a costly, time-consuming, and complicated strategy. So, sgRNA designing could be more effective [108, 109]. Furthermore, an effective sgRNA should simultaneously show the highest on-target efficiency and the lowest off-target activity. So, several well-developed computational tools can be found to design sgRNA for high-efficiency genome editing [110]. In addition to the simplicity, high efficiency, and cost-effectiveness, the in silico tool offers adaptability, automation, and fast processing to analyze many genes [111].

In the last decade, many in silico tools have been introduced to developing CRISPR technology because there has been an urgent demand to design an effective sgRNA to create precise mutations via CRISPR/Cas. Some tools have combined several scoring methods and/or algorithms to provide better design services [112, 113]. In addition to the different features of effectively designing sgRNA, these tools would be user-friendly [114]. The most popular computational sgRNA designing tools are outlined in the following and the other in silico tools are summarized in Table 4. The outlined tools are able to offer candidate sgRNAs and simultaneously score off-target activity. Besides, they are more user-friendly and fast-processing tools [28, 44].

4.1 CHOPCHOP

CHOPCHOP website, one of the most conventional in silico tools for detecting target sequences, includes a clear interface and complete functions. There are more than 200 reference genomes on the website so that the users can search for the target sequence, genomic coordinates, and name of the desired gene. The users are also able to choose two different methods to detect off-target and seven scoring approaches for on-target efficiency before sgRNA designing (Table 5) [117, 125, 126].

StrategiesAdvantagesLimitationRef.
Cas engineeringPAMs improvement, shortening sgRNAsCostly, time-consuming, reforming for each propose[98]
Transcriptome analysisPrecisely detecting on- and off-target activityCostly, time-consuming[99]
Tunable systemRegulating Cas 9 working time, reducing undesirable DNA cleavageSlow on-rate, decreased editing efficiency, time-consuming for inducing by light or chemicals[100]
Functional screeningValidation of gene functions, controlling genetic disruptionJust applicable for the low-throughput formats[101]
RNP complexReduced off-target effects, transient genome editingHigh molecular weight, inducing phospholipid bilayer stress[102]
Truncated sgRNAMinimizing off-target effects without reducing on-target activity, decreasing sgRNA lengthEditing in some new off-target sequences[103]
Paired Cas9 nickaseHigh efficiency and specificity, using for insertion in target gene by NHEJRequiring two simultaneous sgRNAs[104]

Table 3.

Benefits and limitations of the strategies predicting off-target effects.

NameServicesURLRef.
sgRNA DesignersgRNA designinghttps://portals.broadinstitute.org/gpp/public/analysis-tools/sgrna-design[115]
CRISPR-ERAsgRNA designinghttp://CRISPR-ERA.stanford.edu[116]
CHOPCHOPDetecting target sequencehttps://chopchop.cbu.uib.no/about[117]
CRISPRscansgRNA designing and analysishttps://www.crisprscan.org/[118]
CRISPR-GEsgRNA designing detecting target sequencehttp://skl.scau.edu.cn/[119]
CRISPR RGEN ToolsDetecting off-target siteshttp://www.rgenome.net/[120]
sgRNAcas9sgRNA designing predicting off-target siteshttp://biootools.com/[121]
CRISPR MultiTargeterMultiple sgRNA designing toolhttp://multicrispr.net/[122]
CRISPR-PsgRNA designing for plantshttp://crispr.hzau.edu.cn/CRISPR2/[123]
CRISPORsgRNA designing
Detecting on- and off-target sites
http://crispor.tefor.net/[124]

Table 4.

Common in silico tools for sgRNA designing in CRISPR technology.

ToolCovered genomeNucleasesNickaseOff-target analysisCas type
ChopChop>200YesNoYesDifferent Type II
CRISPR RGEN350YesNoYesDifferent Type II
CRISPOR417YesNoYesDifferent Type II

Table 5.

Comparing three in silico tools.

4.2 CRISPR RGEN tools

CRISPR/Cas-derived RNA-guided engineered nuclease (CRISPR RGEN) can provide several computational tools and sgRNA/Cas libraries, including nine tools such as Cas-OFFinder, Cas-Designer, and Digenome-Seq. Compared with other tools, Cas-Designer can rapidly detect potential off-target sites containing a DNA or RNA bulge. Besides, Cas-Designer can offer an out-of-frame score for each sgRNA to find the proper sites for the gene knockout [120, 127, 128]. Cas-OFFinder is also used to seek potential off-target positions of NmCas9 endonuclease (from N.meningitides), recognizing 5′-NNNNGMTT-3′ PAM sequence (M = A or C) as well as a 24-bp target sequence specific to the design sgRNA in the target genome. Also, mixed bases can be used by Cas-OFFinder to analyze the degeneracy of PAM sequences. At least, Cas-OFFinder can provide quick scanning for potential off-target positions in any sequenced genome, regardless of the mismatched nucleotides numbers or the PAM sequence limitation (Table 5) [120, 129, 130].

4.3 CRISPOR

Among these in silico tools for effectively designing sgRNA, CRISPOR includes various useful tools to design sgRNA, 417 genomes, and 19 PAM types. This in silico tool can receive genome coordinates and sequences with more than 2000 bp length as the inputs. After processing, comprehensive information is provided as the output. The result can be, by default, presented in two sections; first, visualizing the PAM sites along the target sequence, available in different formats such as fasta, GenBank and SnapGene; second, providing a table containing all information such as 2 specificity scores and 10 efficiency scores for every predicted sgRNA (Table 5) [105, 124, 131, 132].

4.4 Challenges and limitations of in silico tools

Although the computational tools for sgRNA designing can facilitate on-target prediction and reduce off-target activity, some tools cannot cover all vital criteria. So, it is highly recommended to use different in silico tools. For example, some free, user-friendly, and reliable online tools include RNAfold and Mfold to predict sgRNA secondary structure [59, 79].

Moreover, all current prediction models struggle with four main challenges:

  1. Data insufficiency: machine learning models (MLDs) operate better than other approaches due to their data-driven process. Nevertheless, they cannot accurately anticipate previously unseen data without sufficient data to fully extract features [133, 134].

  2. Unclear mechanism: The mechanism of the CRISPR/Cas9 editing system has not been completely discovered and limits the characteristics used in the most recent cutting-edge algorithms. So, MDL-mediated approaches are not simply able to achieve significant advancements with sufficient data [133, 134].

  3. Data heterogeneity: generated datasets from various platforms and cell types should be merged for data augmentation [133].

  4. Data imbalance: the detection of the frequency of off-target sequences through high-throughput whole-genome sequencing is considerably less than the numbers detected by in silico prediction tools [133, 135].

Advertisement

5. Adenine base editors (ABEs)

Base editors (BEs), as chimeric proteins, contain a catalytic domain and a DNA targeting modules which is able to deaminate adenines and cytosines. There is no need to make DSBs in DNA bases editing when BEs are used for base editing. So, these proteins can reduce the off-target effects and random indels at the on-target sequences. The BEs have been introduced as novel promising tools to make precise gene modification [6, 136, 137]. ABEs are the fused Cas9 nickase with a deaminase domain converting A-G and C-T (C > T) at the target sequence [138]. In fact, ABEs can effectively and precisely convert A-G and C-T base pairs at the target site within the editing frame while producing few by-products, consequently reducing off-target activity significantly. The ABE variants can improve the precision of adenine base editing by reducing the off-target activities of RNA and DNA [105, 139]. It has been reported that ABEs cooperate to induce free-sgRNA transcriptome editing. ABEs have also been discovered to display RNA off-target activity and the capacity to edit their own transcripts [140, 141]. ABEs can generally produce significant off-target single-nucleotide variations (SNVs) in RNA sequences. Therefore, deaminase engineering enables ABE variants to decrease off-target mutation of SNVs in RNA sequences while increasing on-target efficiency with DNA [105, 142].

Advertisement

6. Conclusion

Over the last 30 years, genome editing technology, particularly CRISPR/Cas, has promoted biosciences by editing and targeting the genomic DNAs of any species. CRISPR/Cas is the most precise, effective, and affordable among all these genome editing technologies. Although there are diverse types and classes of CRISPR/Cas systems, they are not all applicable due to the high rate of off-target effects. Different approaches have been developed to decrease the off-target effects for enhancing the precision and efficiency of the different CRISPR/Cas techniques. According to a refined reference genome, a well-designed sgRNA can support high on-target efficiency to create a precise and desirable mutation. Finally, it is highly recommended to consider the criteria as mentioned above, including GC content, length, secondary structure, and sequence, for designing an effective sgRNA to achieve high-precision CRISPR genome editing.

Advertisement

Acknowledgments

The chapter was funded by Amin Techno Gene Private Virtual Lab (NGO), Tehran, Iran.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Dhawan M, Sharma M, Grewal RS. CRISPR Systems: RNA-Guided defence mechanisms in Bacteria and Archaea. International Journal of Current Microbiology and Applied Sciences. 2015;4(6):187-200
  2. 2. Makarova KS, Wolf YI, Iranzo J, Shmakov SA, Alkhnbashi OS, Brouns SJ, et al. Evolutionary classification of CRISPR–Cas systems: A burst of class 2 and derived variants. Nature Reviews Microbiology. 2020;18(2):67-83
  3. 3. Koonin EV, Makarova KS, Wolf YI. Evolutionary genomics of defense systems in archaea and bacteria. Annual Review of Microbiology. 2017;71:233-261
  4. 4. Adli M. The CRISPR tool kit for genome editing and beyond. Nature Communications. 2018;9(1):1-13
  5. 5. Pickar-Oliver A, Gersbach CA. The next generation of CRISPR–Cas technologies and applications. Nature Reviews Molecular Cell Biology. 2019;20(8):490-507
  6. 6. Komor AC, Kim YB, Packer MS, Zuris JA, Liu DR. Programmable editing of a target base in genomic DNA without double-stranded DNA cleavage. Nature. 2016;533(7603):420-424
  7. 7. Hille F, Richter H, Wong SP, Bratovič M, Ressel S, Charpentier E. The biology of CRISPR-Cas: Backward and forward. Cell. 2018;172(6):1239-1259
  8. 8. Min YL, Bassel-Duby R, Olson EN. CRISPR correction of Duchenne muscular dystrophy. Annual Review of Medicine. 2019;70:239-255
  9. 9. Xu CF, Chen GJ, Luo YL, Zhang Y, Zhao G, Lu ZD, et al. Rational designs of in vivo CRISPR-Cas delivery systems. Advanced Drug Delivery Reviews. 2021;168:3-29
  10. 10. Mohammadhassan R, Akhavan S, Mahmoudi A, Khalkhali A, Barzin R. Antiviral activity of Echinacea (Echinacea Purpurea). International Journal of Advanced Biotechnology and Research. 2016;7(4):1493-1497
  11. 11. Hynes AP, Rousseau GM, Agudelo D, Goulet A, Amigues B, Loehr J, et al. Widespread anti-CRISPR proteins in virulent bacteriophages inhibit a range of Cas9 proteins. Nature Communications. 2018;9(1):1-10
  12. 12. Doudna JA, Charpentier E. The new frontier of genome engineering with CRISPR-Cas9. Science. 2014;2014:346
  13. 13. Hess GT, Tycko J, Yao D, Bassik MC. Methods and applications of CRISPR-mediated base editing in eukaryotic genomes. Molecular Cell. 2017;68(1):26-43
  14. 14. Ran FA, Cong L, Yan WX, Scott DA, Gootenberg JS, Kriz AJ, et al. In vivo genome editing using Staphylococcus aureus Cas9. Nature. 2015;520(7546):186-191
  15. 15. Wright AV, Nuñez JK, Doudna JA. Biology and applications of CRISPR systems: Harnessing nature’s toolbox for genome engineering. Cell. 2016;164(1-2):29-44
  16. 16. Leenay RT. Deciphering, communicating, and engineering the CRISPR PAM. Journal of Molecular Biology. 2017;429(2):177-191
  17. 17. Meeske AJ. RNA guide complementarity prevents self-targeting in type VI CRISPR systems. Molecular Cell. 2018;71(5):791-801
  18. 18. Liu Y, Pinto F, Wan X, Peng S, Li M, Xie Z, et al. Reprogrammed tracrRNAs enable repurposing RNAs as crRNAs and detecting RNAs. bioRxiv. 2018;2021:1-28
  19. 19. Liu Y, Pinto F, Wan X, Yang Z, Peng S, Li M, et al. Reprogrammed tracrRNAs enable repurposing of RNAs as crRNAs and sequence-specific RNA biosensors. Nature Communications. 2022;13(1):1-12
  20. 20. Charpentier E, Richter H, van der Oost J, White MF. Biogenesis pathways of RNA guides in archaeal and bacterial CRISPR-Cas adaptive immunity. FEMS Microbiology Reviews. 2015;39(3):428-441
  21. 21. Jiao C, Sharma S, Dugar G, Peeck NL, Bischler T, Wimmer F, et al. Noncanonical crRNAs derived from host transcripts enable multiplexable RNA detection by Cas9. Science. 2021;372(6545):941-948
  22. 22. Reis AC, Halper SM, Vezeau GE, Cetnar DP, Hossain A, Clauer PR, et al. Simultaneous repression of multiple bacterial genes using nonrepetitive extra-long sgRNA arrays. Nature Biotechnology. 2019;37(11):1294-1301
  23. 23. Nelles DA, Fang MY, O’Connell MR, Xu JL, Markmiller SJ, Doudna JA, et al. Programmable RNA tracking in live cells with CRISPR/Cas9. Cell. 2016;165(3):488-496
  24. 24. Hirano H, Gootenberg JS, Horii T, Abudayyeh OO, Kimura M, Hsu PD, et al. Structure and engineering of Francisella novicida Cas9. Cell. 2016;164(5):950-961
  25. 25. Gasiunas G, Young JK, Karvelis T, Kazlauskas D, Urbaitis T, Jasnauskaite M, et al. A catalogue of biochemically diverse CRISPR-Cas9 orthologs. Nature Communications. 2020;11(1):1-10
  26. 26. Chyou TY. Prediction and diversity of tracrRNAs from type II CRISPR-Cas systems. RNA Biology. 2019;16(3):423-434
  27. 27. Hiranniramol K, Chen Y, Liu W, Wang X. Generalizable sgRNA design for improved CRISPR/Cas9 editing efficiency. Bioinformatics. 2020;36(9):2684-2689
  28. 28. Baghini SS, Gardanova ZR, Zekiy AO, Shomali N, Tosan F, Jarahian M. Optimizing sgRNA to improve CRISPR/Cas9 knockout efficiency: Special focus on human and animal cell. Frontiers in Bioengineering and Biotechnology. 2021;2021:9
  29. 29. Mohanta TK, Bashir T, Hashem A. Genome editing tools in plants. Genes. 2017;8(12):399
  30. 30. Van de Wiel CCM, Schaart JG, Lots LAP, Smulders MJM. New traits in crops produced by genome editing techniques based on deletions. Plant Biotechnology Reports. 2017;11(1):1-8
  31. 31. Mishra R, Zhao K. Genome editing technologies and their applications in crop improvement. Plant Biotechnology Reports. 2018;12(2):57-68
  32. 32. Vats S, Kumawat S, Kumar V, Patil GB, Joshi T, Sonah H, et al. Genome editing in plants: Exploration of technological advancements and challenges. Cell. 2019;8:11
  33. 33. Knott GJ, Doudna JA. CRISPR-Cas guides the future of genetic engineering. Science. 2018;361(6405):866-869
  34. 34. Yang H, Wu JJ, Tang T, Liu KD, Dai C. CRISPR/Cas9-mediated genome editing efficiently creates specific mutations at multiple loci using one sgRNA in Brassica napus. Scientific Reports. 2017;7(1):1-13
  35. 35. Musunuru K. The hope and hype of CRISPR-Cas9 genome editing: A review. JAMA Cardiology. 2017;2(8):914-919
  36. 36. Raper AT, Stephenson AA, Suo Z. Functional insights revealed by the kinetic mechanism of CRISPR/Cas9. Journal of the American Chemical Society. 2018;140(8):2971-2984
  37. 37. Ma X, Zhang Q, Zhu Q, Liu W, Chen Y, Qiu R, et al. A robust CRISPR/Cas9 system for convenient, high-efficiency multiplex genome editing in monocot and dicot plants. Molecular Plant. 2015;8(8):1274-1284
  38. 38. Nadakuduti SS, Enciso-Rodríguez F. Advances in genome editing with CRISPR systems and transformation technologies for plant DNA manipulation. Frontiers in Plant Science. 2021;11:637159
  39. 39. Kang JG, Park JS, Ko JH, Kim YS. Regulation of gene expression by altered promoter methylation using a CRISPR/Cas9-mediated epigenetic editing system. Scientific Reports. 2019;9(1):1-12
  40. 40. Hao M, Cui Y, Qu X. Analysis of CRISPR-Cas system in Streptococcus thermophilus and its application. Frontiers in Microbiology. 2018;9:257
  41. 41. Miller SM, Wang T, Randolph PB, Arbab M, Shen MW, Huang TP, et al. Continuous evolution of SpCas9 variants compatible with non-G PAMs. Nature Biotechnology. 2020;38(4):471-481
  42. 42. Burgess SM. Genome editing by targeted nucleases and the CRISPR/Cas revolution. The Liver: Biology and Pathobiology. 2020;2020:953-964
  43. 43. Fallahi S, Mohammadhassan R. A review of pharmaceutical recombinant proteins and gene transformation approaches in transgenic poultry. Journal of Tropical Life Science. 2020;10(2):163-173
  44. 44. Liu G. Computational approaches for effective CRISPR guide RNA design and evaluation. Computational and Structural Biotechnology Journal. 2020;18:35-44
  45. 45. Rose JC, Popp NA, Richardson CD, Stephany JJ, Mathieu J, Wei CT, et al. Suppression of unwanted CRISPR-Cas9 editing by co-administration of catalytically inactivating truncated guide RNAs. Nature Communications. 2020;11(1):1-11
  46. 46. Xie K, Yang Y. A multiplexed CRISPR/Cas9 editing system based on the endogenous tRNA processing. In: Plant Genome Editing with CRISPR Systems. New York: Humana; 2019. pp. 63-73
  47. 47. Hui L, Zhao M, He J, Hu Y, Huo Y, Hao H, et al. A simple and reliable method for creating PCR-detectable mutants in Arabidopsis with the polycistronic tRNA–gRNA CRISPR/Cas9 system. Acta Physiologiae Plantarum. 2019;41(10):1-14
  48. 48. Vakulskas CA, Behlke MA. Evaluation and reduction of CRISPR off-target cleavage events. Nucleic Acid Therapeutics. 2019;29(4):167-174
  49. 49. Kempton HR, Qi LS. When genome editing goes off-target. Science. 2019;364(6437):234-236
  50. 50. Graham N, Patil GB, Bubeck DM, Dobert RC, Glenn KC, Gutsche AT, et al. Plant genome editing and the relevance of off-target changes. Plant Physiology. 2020;183(4):1453-1471
  51. 51. Chen JS, Dagdas YS, Kleinstiver BP, Welch MM, Sousa AA, Harrington LB, et al. Enhanced proofreading governs CRISPR–Cas9 targeting accuracy. Nature. 2017;550(7676):407-410
  52. 52. Fan R, Chai Z, Xing S, Chen K, Qiu F, Chai T, et al. Shortening the sgRNA-DNA interface enables SpCas9 and eSpCas9 (1.1) to nick the target DNA strand. Life Sciences. 2020;63(11):1619-1630
  53. 53. Tsai SQ, Zheng Z, Nguyen NT, Liebers M, Topkar VV, Thapar V, et al. GUIDE-seq enables genome-wide profiling of off-target cleavage by CRISPR-Cas nucleases. Nature Biotechnology. 2015;33(2):187-197
  54. 54. Hajiahmadi Z, Movahedi A, Wei H, Li D, Orooji Y, Ruan H, et al. Strategies to increase on-target and reduce off-target effects of the CRISPR/Cas9 system in plants. International Journal of Molecular Sciences. 2019;20:15
  55. 55. Jensen KT, Fløe L, Petersen TS, Huang J, Xu F, Bolund L, et al. Chromatin accessibility and guide sequence secondary structure affect CRISPR-Cas9 gene editing efficiency. FEBS Letters. 2017;591(13):1892-1901
  56. 56. Wang D, Zhang C, Wang B, Li B, Wang Q, Liu D, et al. Optimized CRISPR guide RNA design for two high-fidelity Cas9 variants by deep learning. Nature Communications. 2019;10(1):1-14
  57. 57. Tatiossian KJ, Clark RD, Huang C, Thornton ME, Grubbs BH, Cannon PM. Rational selection of CRISPR-Cas9 guide RNAs for homology-directed genome editing. Molecular Therapy. 2021;29(3):1057-1069
  58. 58. Dang Y, Jia G, Choi J, Ma H, Anaya E, Ye C, et al. Optimizing sgRNA structure to improve CRISPR-Cas9 knockout efficiency. Genome Biology. 2015;16(1):1-10
  59. 59. Schindele P, Wolter F, Puchta H. CRISPR guide RNA design guidelines for efficient genome editing. In: RNA Tagging. New York: Humana; 2020. pp. 331-342
  60. 60. Dhanjal JK, Dammalapati S, Pal S, Sundar D. Evaluation of off-targets predicted by sgRNA design tools. Genomics. 2020;112(5):3609-3614
  61. 61. Chang AY. Single-guide RNAs: Rationale and design. In: CRISPR Genome Surgery in Stem Cells and Disease Tissues. England: Academic Press; 2022. pp. 47-55
  62. 62. Ren F, Ren C, Zhang Z, Duan W, Lecourieux D, Li S, et al. Efficiency optimization of CRISPR/Cas9-mediated targeted mutagenesis in grape. Frontiers in Plant Science. 2019;10:612
  63. 63. Xu Hua F, Wainberg M, Kundaje A, Fire AZ. High-throughput characterization of Cascade type IE CRISPR guide efficacy reveals unexpected PAM diversity and target sequence preferences. Genetics. 2017;206(4):1727-1738
  64. 64. Bruegmann T, Deecke K, Fladung M. Evaluating the efficiency of gRNAs in CRISPR/Cas9 mediated genome editing in poplars. International Journal of Molecular Sciences. 2019;20:15
  65. 65. Liu Y, Wei WP, Ye BC. High GC content Cas9-mediated genome-editing and biosynthetic gene cluster activation in Saccharopolyspora erythraea. ACS Synthetic Biology. 2018;7(5):1338-1348
  66. 66. Zhao C, Wang Y, Nie X, Han X, Liu H, Li G, et al. Evaluation of the effects of sequence length and microsatellite instability on single-guide RNA activity and specificity. International Journal of Biological Sciences. 2019;15(12):2641
  67. 67. Karmakar S, Behera D, Baig MJ, Molla KA. In vitro Cas9 cleavage assay to check guide RNA efficiency. In: CRISPR-Cas Methods. New York: Humana; 2021. pp. 23-39
  68. 68. Li J, Hong S, Chen W, Zuo E, Yang H. Advances in detecting and reducing off-target effects generated by CRISPR-mediated genome editing. Journal of Genetics and Genomics. 2019;46(11):513-521
  69. 69. Zhu LJ. Overview of guide RNA design tools for CRISPR-Cas9 genome editing technology. Frontiers in Biology. 2015;10(4):289-296
  70. 70. Lv J, Wu S, Wei R, Li Y, Jin J, Mu Y, et al. The length of guide RNA and target DNA heteroduplex effects on CRISPR/Cas9 mediated genome editing efficiency in porcine cells. Journal of Veterinary Science. 2019;20:3
  71. 71. Zhang JP, Li XL, Neises A, Chen W, Hu LP, Ji GZ, et al. Different effects of sgRNA length on CRISPR-mediated gene knockout efficiency. Scientific Reports. 2016;6(1):1-10
  72. 72. Matson AW, Hosny N, Swanson ZA, Hering BJ, Burlak C. Optimizing sgRNA length to improve target specificity and efficiency for the GGTA1 gene using the CRISPR/Cas9 gene editing system. PLoS One. 2019;14:12
  73. 73. Xu J, Lian W, Jia Y, Li L, Huang Z. Optimized guide RNA structure for genome editing via Cas9. Oncotarget. 2017;8:55
  74. 74. Hassan MM, Chowdhury AK, Islam T. In silico analysis of gRNA secondary structure to predict its efficacy for plant genome editing. In: CRISPR-Cas Methods. New York: Humana; 2021. pp. 15-22
  75. 75. Liang G, Zhang H, Lou D, Yu D. Selection of highly efficient sgRNAs for CRISPR/Cas9-based plant genome editing. Scientific Reports. 2016;6(1):1-8
  76. 76. Uniyal AP, Mansotra K, Yadav SK, Kumar V. An overview of designing and selection of sgRNAs for precise genome editing by the CRISPR-Cas9 system in plants. 3 Biotech. 2019;9(6):1-9
  77. 77. Liang Y, Eudes A, Yogiswara S, Jing B, Benites VT, Yamanaka R, et al. A screening method to identify efficient sgRNAs in Arabidopsis, used in conjunction with cell-specific lignin reduction. Biotechnology for Biofuels. 2019;12(1):1-15
  78. 78. Jiang M, Ye Y, Li J. Core Hairpin Structure of SpCas9 sgRNA functions in a sequence-and spatial conformation–dependent manner. Translating Life Sciences Innovation. 2021;26(1):92-102
  79. 79. Dong C, Gou Y, Lian J. SgRNA engineering for improved genome editing and expanded functional assays. Current Opinion in Biotechnology. 2022;75:102697
  80. 80. Kocak DD, Josephs EA, Bhandarkar V, Adkar SS, Kwon JB, Gersbach CA. Increasing the specificity of CRISPR systems with engineered RNA secondary structures. Nature Biotechnology. 2019;37(6):657-666
  81. 81. Hu Z, Wang Y, Liu Q, Qiu Y, Zhong Z, Li K, et al. Improving the precision of base editing by bubble hairpin single guide RNA. MBio. 2021;12:2
  82. 82. Konstantakos V, Nentidis A, Krithara A, Paliouras G. CRISPR–Cas9 gRNA efficiency prediction: An overview of predictive tools and the role of deep learning. Nucleic Acids Research. 2022;50(7):3616-3637
  83. 83. Thyme SB, Akhmetova L, Montague TG, Valen E, Schier AF. Internal guide RNA interactions interfere with Cas9-mediated cleavage. Nature Communications. 2016;7(1):1-7
  84. 84. Muhammad Rafid AH, Toufikuzzaman M, Rahman MS, Rahman MS. CRISPRpred (SEQ): A sequence-based method for sgRNA on target activity prediction using traditional machine learning. BMC Bioinformatics. 2020;21(1):1-13
  85. 85. Wong N, Liu W, Wang X. WU-CRISPR: Characteristics of functional guide RNAs for the CRISPR/Cas9 system. Genome Biology. 2015;16(1):1-8
  86. 86. Wilson LO, O’Brien AR, Bauer DC. The current state and future of CRISPR-Cas9 gRNA design tools. Frontiers in Pharmacology. 2018;9:749
  87. 87. Gerashchenkov GA, Rozhnova NA, Kuluev BR, Kiryanova OY, Gumerova GR, Knyazev AV, et al. Design of guide RNA for CRISPR/Cas plant genome editing. Molecular Biology. 2020;54(1):24-42
  88. 88. Horodecka K, Düchler M. CRISPR/Cas9: Principle, applications, and delivery through extracellular vesicles. International Journal of Molecular Sciences. 2021;22:11
  89. 89. Hanna RE. Design and analysis of CRISPR–Cas experiments. Nature Biotechnology. 2020;38(7):813-823
  90. 90. Kim N, Kim HK, Lee S, Seo JH, Choi JW, Park J, et al. Prediction of the sequence-specific cleavage activity of Cas9 variants. Nature Biotechnology. 2020;38(11):1328-1336
  91. 91. Moreb EA, Lynch MD. Genome dependent Cas9/gRNA search time underlies sequence dependent gRNA activity. Nature Communications. 2021;12(1):1-13
  92. 92. Xiang X, Corsi GI, Anthon C, Qu K, Pan X, Liang X, et al. Enhancing CRISPR-Cas9 gRNA efficiency prediction by data integration and deep learning. Nature Communications. 2021;12(1):1-9
  93. 93. Corsi GI, Qu K, Alkan F, Pan X, Luo Y, Gorodkin J. CRISPR/Cas9 gRNA activity depends on free energy changes and on the target PAM context. Nature Communications. 2022;13(1):1-14
  94. 94. Sledzinski P, Nowaczyk M, Olejniczak M. Computational tools and resources supporting CRISPR-Cas experiments. Cell. 2020;9:5
  95. 95. Li C, Chu W, Gill RA, Sang S, Shi Y, Hu X, et al. Computational tools and resources for CRISPR/Cas genome editing. Genomics, Proteomics and Bioinformatics. 2022 (In Press). DOI: 10.1016/j.gpb.2022.02.006
  96. 96. Luo M, Wang J, Dong Z, Wang C, Lu G. CRISPR-Cas9 sgRNA design and outcome assessment: Bioinformatics tools and aquaculture applications. Aquaculture and Fisheries. 2021;7(2):121-130
  97. 97. Aslam MA, Hammad M, Ahmad A, Altenbuchner J, Ali H. Delivery methods, resources and design tools in CRISPR/Cas. In: CRISPR Crops. Singapore: Springer; 2021. pp. 63-116
  98. 98. Slaymaker IM, Gaudelli NM. Engineering Cas9 for human genome editing. Current Opinion in Structural Biology. 2021;69:86-98
  99. 99. Hrdlickova R, Toloue M, Tian B. RNA-Seq methods for transcriptome analysis. Wiley Interdisciplinary Reviews: RNA. 2017;8:1
  100. 100. Nuñez JK, Harrington LB, Doudna JA. Chemical and biophysical modulation of Cas9 for tunable genome engineering. ACS Chemical Biology. 2016;11(3):681-688
  101. 101. Sun N, Petiwala S, Wang R, Lu C, Hu M, Ghosh S, et al. Development of drug-inducible CRISPR-Cas9 systems for large-scale functional screening. BMC Genomics. 2019;20(1):1-15
  102. 102. Zhang S, Shen J, Li D, Cheng Y. Strategies in the delivery of Cas9 ribonucleoprotein for CRISPR/Cas9 genome editing. Theranostics. 2021;11(2):614
  103. 103. Cui Z, Tian R, Huang Z, Jin Z, Li L, Liu J, et al. FrCas9 is a CRISPR/Cas9 system with high editing efficiency and fidelity. Nature Communications. 2022;13(1):1-12
  104. 104. Gopalappa R, Suresh B, Ramakrishna S, Kim H. Paired D10A Cas9 nickases are sometimes more efficient than individual nucleases for gene disruption. Nucleic Acids Research. 2018;46(12):e71
  105. 105. Manghwar H, Li B, Ding X, Hussain A, Lindsey K, Zhang X, et al. CRISPR/Cas systems in genome editing: Methodologies and tools for sgRNA design, off-target evaluation, and strategies to mitigate off-target effects. Advanced Science. 2020;7(6):1902312
  106. 106. Asmamaw M, Zawdie B. Mechanism and applications of CRISPR/Cas-9-mediated genome editing. Biologics: Targets and Therapy. 2021;15:353
  107. 107. Nasir MA, Nawaz S, Huang J. A review: Computational approaches to design sgRNA of CRISPR-Cas9. Current Bioinformatics. 2022;17(1):2-18
  108. 108. Xue L, Tang B, Chen W, Luo J. Prediction of CRISPR sgRNA activity using a deep convolutional neural network. Journal of Chemical Information and Modeling. 2018;59(1):615-624
  109. 109. Xiao LM, Wan YQ, Jiang ZR. AttCRISPR: A spacetime interpretable model for prediction of sgRNA on-target activity. BMC Bioinformatics. 2020;22(1):1-17
  110. 110. Chuai GH, Wang QL, Liu Q. In silico meets in vivo: Towards computational CRISPR-based sgRNA design. Trends in Biotechnology. 2017;35(1):12-21
  111. 111. Karlapudi AP, Venkateswarulu TC, Tammineedi J, Srirama K, Kanumuri L, Kodali VP. In silico sgRNA tool design for CRISPR control of quorum sensing in Acinetobacter species. Genes and Diseases. 2018;5(2):123-129
  112. 112. Choudhary S, Ubale A, Padiya J, Mikkilineni V. Application of bioinformatics tools in CRISPR/Cas. In: CRISPR/Cas Genome Editing. Cham: Springer; 2020. pp. 31-52
  113. 113. Peng H, Zheng Y, Zhao Z, Li J. Multigene editing: Current approaches and beyond. Briefings in Bioinformatics. 2021;22:5
  114. 114. Zhou H, Zhou M, Li D, Manthey J, Lioutikova E, Wang H, et al. Whole genome analysis of CRISPR Cas9 sgRNA off-target homologies via an efficient computational algorithm. BMC Genomics. 2017;18(9):31-38
  115. 115. Doench JG, Fusi N, Sullender M, Hegde M, Vaimberg EW, Donovan KF, et al. Optimized sgRNA design to maximize activity and minimize off-target effects of CRISPR-Cas9. Nature Biotechnology. 2016;34(2):184-191
  116. 116. Liu H, Wei Z, Dominguez A, Li Y, Wang X, Qi LS. CRISPR-ERA: A comprehensive design tool for CRISPR-mediated gene editing, repression and activation. Bioinformatics. 2015;31(22):3676-3678
  117. 117. Montague TG, Cruz JM, Gagnon JA, Church GM, Valen E. CHOPCHOP: A CRISPR/Cas9 and TALEN web tool for genome editing. Nucleic Acids Research. 2014;42(W1):W401-W407
  118. 118. Moreno-Mateos MA, Vejnar CE, Beaudoin JD, Fernandez JP, Mis EK, Khokha MK, et al. CRISPRscan: Designing highly efficient sgRNAs for CRISPR-Cas9 targeting in vivo. Nature Methods. 2015;12(10):982-988
  119. 119. Xie X, Ma X, Zhu Q, Zeng D, Li G, Liu YG. CRISPR-GE: A convenient software toolkit for CRISPR-based genome editing. Molecular Plant. 2017;10(9):1246-1249
  120. 120. Hwang G-H, Kim J-S, Bae S. Web-based CRISPR toolkits: Cas-OFFinder, cas-designer, and cas-analyzer. In: CRISPR Guide RNA Design. New York, NY: Springer; 2021. pp. 23-33
  121. 121. Xie S, Shen B, Zhang C, Huang X, Zhang Y. sgRNAcas9: A software package for designing CRISPR sgRNA and evaluating potential off-target cleavage sites. PloS One. 2014;9(6):e100448
  122. 122. Prykhozhij SV, Rajan V, Gaston D, Berman JN. CRISPR multitargeter: A web tool to find common and unique CRISPR single guide RNA targets in a set of similar sequences. PLoS One. 2015;10(3):e0119372
  123. 123. Lei Y, Lu L, Liu HY, Li S, Xing F, Chen LL. CRISPR-P: A web tool for synthetic single-guide RNA design of CRISPR-system in plants. Molecular Plant. 2014;7(9):1494-1496
  124. 124. Concordet J-P, Haeussler M. CRISPOR: Intuitive guide selection for CRISPR/Cas9 genome editing experiments and screens. Nucleic Acids Research. 2018;46(W1):W242-W245
  125. 125. Labun K, Montague TG, Gagnon JA, Thyme SB, Valen E. CHOPCHOP v2: A web tool for the next generation of CRISPR genome engineering. Nucleic Acids Research. 2016;44(W1):W272-W276
  126. 126. Labun K, Krause M, Torres Cleuren Y, Valen E. CRISPR genome editing made easy through the CHOPCHOP website. Current Protocols. 2021;1(4):e46
  127. 127. Park J, Bae S, Kim J-S. Cas-Designer: A web-based tool for choice of CRISPR-Cas9 target sites. Bioinformatics. 2015;31(24):4014-4016
  128. 128. Bradford J, Perrin D. A benchmark of computational CRISPR-Cas9 guide design methods. PLoS Computational Biology. 2019;15(8):e1007274
  129. 129. Bae S, Park J, Kim J-S. Cas-OFFinder: A fast and versatile algorithm that searches for potential off-target sites of Cas9 RNA-guided endonucleases. Bioinformatics. 2014;30(10):1473-1475
  130. 130. Zhao C et al. CRISPR-offinder: A CRISPR guide RNA design and off-target searching tool for user-defined protospacer adjacent motif. International Journal of Biological Sciences. 2017;13(12):1470
  131. 131. Haeussler M et al. Evaluation of off-target and on-target scoring algorithms and integration into the guide RNA selection tool CRISPOR. Genome Biology. 2016;17(1):1-12
  132. 132. Haeussler M. CRISPR off-targets: A question of context. Cell Biology and Toxicology. 2020;36(1):5-9
  133. 133. Wang J, Zhang X, Cheng L, Luo Y. An overview and metanalysis of machine and deep learning-based CRISPR gRNA design tools. RNA Biology. 2020;17(1):13-22
  134. 134. Jin J, Huo L, Song S, Ajani JA. CRISPR/Cas9 Technology Improvements and the RNA-Editing Trend. Journal of Molecular Biology and Biotechnology. 2020;5:5
  135. 135. Li W, Wang XB, Xu Y. Recognition of CRISPR Off-target cleavage sites with SeqGAN. Current Bioinformatics. 2022;17(1):101-107
  136. 136. Kim D, Kim DE, Lee G, Cho SI, Kim JS. Genome-wide target specificity of CRISPR RNA-guided adenine base editors. Nature Biotechnology. 2019;37(4):430-435
  137. 137. Jin S, Zong Y, Gao Q, Zhu Z, Wang Y, Qin P, et al. Cytosine, but not adenine, base editors induce genome-wide off-target mutations in rice. Science. 2019;364(6437):292-295
  138. 138. Gaudelli NM, Komor AC, Rees HA, Packer MS, Badran AH, Bryson DI, et al. Programmable base editing of A• T to G• C in genomic DNA without DNA cleavage. Nature Biotechnology. 2017;551(7681):464-471
  139. 139. Rees HA, Wilson C, Doman JL, Liu DR. Analysis and minimization of cellular RNA editing by DNA adenine base editors. Science Advances. 2019;5:5
  140. 140. Grünewald J, Zhou R, Garcia SP, Iyer S, Lareau CA, Aryee MJ, et al. Transcriptome-wide off-target RNA editing induced by CRISPR-guided DNA base editors. Nature. 2019;569(7756):433-437
  141. 141. Grünewald J, Zhou R, Iyer S, Lareau CA, Garcia SP, Aryee MJ, et al. CRISPR DNA base editors with reduced RNA off-target and self-editing activities. Nature Biotechnology. 2019;37(9):1041-1048
  142. 142. Zhou C, Sun Y, Yan R, Liu Y, Zuo E, Gu C, et al. Off-target RNA mutation induced by DNA base editing and its elimination by mutagenesis. Nature. 2019;571(7764):275-278

Written By

Reza Mohammadhassan, Sara Tutunchi, Negar Nasehi, Fatemeh Goudarziasl and Lena Mahya

Submitted: 25 June 2022 Reviewed: 22 July 2022 Published: 19 August 2022