Link to Publisher's site
Plant Physiol. 2017 Aug; 174(4): 2023–2037.
Published online 2017 Jun 23. doi: 10.1104/pp.17.00489
PMCID: PMC5543939
PMID: 28646085

Generation of a Collection of Mutant Tomato Lines Using Pooled CRISPR Libraries1

Abstract

The high efficiency of clustered regularly interspaced short palindromic repeats (CRISPR)-mediated mutagenesis in plants enables the development of high-throughput mutagenesis strategies. By transforming pooled CRISPR libraries into tomato (Solanum lycopersicum), collections of mutant lines were generated with minimal transformation attempts and in a relatively short period of time. Identification of the targeted gene(s) was easily determined by sequencing the incorporated guide RNA(s) in the primary transgenic events. From a single transformation with a CRISPR library targeting the immunity-associated leucine-rich repeat subfamily XII genes, heritable mutations were recovered in 15 of the 54 genes targeted. To increase throughput, a second CRISPR library was made containing three guide RNAs per construct to target 18 putative transporter genes. This resulted in stable mutations in 15 of the 18 targeted genes, with some primary transgenic plants having as many as five mutated genes. Furthermore, the redundancy in this collection of plants allowed for the association of aberrant T0 phenotypes with the underlying targeted genes. Plants with mutations in a homolog of an Arabidopsis (Arabidopsis thaliana) boron efflux transporter displayed boron deficiency phenotypes. The strategy described here provides a technically simple yet high-throughput approach for generating a collection of lines with targeted mutations and should be applicable to any plant transformation system.

Functional genomics depends on the availability of gene-silencing approaches or mutagenized populations. In plants, the Arabidopsis (Arabidopsis thaliana) T-DNA insertion lines (Berardini et al., 2015), as well as transposon collections in maize (Zea mays; McCarty et al., 2005) and Medicago truncatula (Tadege et al., 2008) and TILLING populations in wheat (Triticum aestivum; http://www.wheat-tilling.com), are powerful resources, provided an insertion or point mutation has been identified in the gene(s) of interest. In tomato (Solanum lycopersicum), several mutagenized populations are available derived from fast-neutron, ethyl methanesulfonate (EMS), TILLING, and transposon-tagging approaches (Menda et al., 2004; Minoia et al., 2010; Saito et al., 2011; Carter et al., 2013). In cases where mutants are not readily available, researchers may rely on transitive or stable gene-silencing approaches to study genes of interest (Smith et al., 2000; Burch-Smith et al., 2004). The ability to make targeted knockout lines with clustered regularly interspaced short palindromic repeats (CRISPR) systems now provides a particularly powerful alternative approach for plant functional genomics studies.

The standard CRISPR system used for targeted mutagenesis is composed of two parts: a CRISPR-associated9 (Cas9) endonuclease and a guide RNA (gRNA) molecule (Jinek et al., 2012). Cas9 identifies a target DNA by base pairing between the target and the gRNA. For the Streptococcus pyogenes Cas9, the protospacer adjacent motif (PAM) is a three-nucleotide NGG motif required for binding to the target DNA sequence. In systems where the gRNA is expressed by the host cell, the gRNA is generally expressed from a Pol III promoter such as U6 or U3, which require an initial G or A, respectively (Bortesi and Fischer, 2015). Upon binding of the target DNA by the Cas9:gRNA complex, Cas9 initiates a double-stranded break (DSB) between the third and fourth nucleotides, 5′ to the PAM (Jinek et al., 2012). Repair of the DSB by the error-prone nonhomologous end-joining pathway typically results in short insertions and/or deletions that can produce knockout mutations in protein-coding sequences (Mladenov and Iliakis, 2011).

Numerous plant species have been modified with CRISPR technology, often with a high rate of mutagenesis. Generally, ∼10% of cells are modified in transient systems, and higher levels of DNA modification are achieved via stable integration of the CRISPR components (Bortesi and Fischer, 2015). However, there is a wide range of efficiencies reported in the literature that may be attributed to different species, reagent delivery, and selection methods. In tomato, CRISPR routinely produces a high frequency of mutagenesis, with biallelic mutants sometimes being observed in the T0 generation (primary transformants; Brooks et al., 2014).

The high mutagenesis frequencies and easy-to-design targeting schemes have allowed for the development of genome-wide CRISPR screens, which were first reported in animal cell cultures (Shalem et al., 2014; Wang et al., 2014). The delivery of pooled CRISPR libraries to cell cultures results in populations of mutagenized cells. In combination with high-throughput sequencing, large-scale genetic screens are being increasingly performed in animal cell cultures (Shalem et al., 2015). Such a system has not yet been reported for plant cell cultures.

In plants, cotransformation with Agrobacterium tumefaciens typically results in ∼20% of the events containing both T-DNAs, with the remaining events incorporating one T-DNA or the other (Depicker et al., 1985; De Buck et al., 2009). Cotransformation with multiple T-DNAs often has been used to generate events with insertions at unlinked loci, such that one locus (e.g. the selectable marker) can be segregated out in future generations, or to increase the throughput of obtaining a collection of transgenic lines (Lein et al., 2008; Char et al., 2017).

We reasoned that, by using a pool of CRISPR T-DNAs for tomato transformation, it would be possible to recover a collection of targeted knockout mutants (Fig. 1). The targeted gene(s) could be determined by sequencing the incorporated gRNA and would allow for easy identification of mutant alleles. The advantage of such a system would be that numerous mutants could be obtained from a single transformation experiment, increasing the throughput of generating knockout lines that could be used in functional studies.

The concept of using a pooled CRISPR library screen in plants. A pool of CRISPR vectors is cloned, maintained as a library in A. tumefaciens, and transformed en masse into a plant system. The resulting T0 plants will generally contain a single T-DNA insertion and a mutation in a single gene (colored outline). Identification of the target genes can be accomplished by sequencing the incorporated gRNA(s). Some plants will not be mutated (wild type; no outline), and ∼20% of plants will contain multiple mutations due to multiple T-DNA insertions (multicolor outline). The T0 plants will give rise to T1 lines segregating for mutant and wild-type alleles and can be used in functional screens or to establish true-breeding mutant lines for future characterization.

RESULTS

Targeting the LRR-RLK Subfamily XII Gene Family in Tomato

In plants, many receptor-like kinases (RLKs) are involved in the perception of pathogens (Schwessinger and Ronald, 2012). The leucine-rich repeat (LRR) subfamily XII (LRR-XII) in particular contains several known receptors of microbe-associated molecular patterns: FLS2, Fls3, EFR, XA21, and CORE (Song et al., 1995; Gómez-Gómez and Boller, 2000; Zipfel et al., 2006; Hind et al., 2016; Wang et al., 2016). However, relatively little is known about the larger LRR-XII gene family and whether other family members play a role in plant immunity. Previous RNA-seq experiments revealed that 21 of the 55 tomato LRR-XII family members (Zheng et al., 2016) have significant differential expression (P ≤ 0.05; 18 induced and three suppressed) during pattern-triggered immunity (PTI; Rosli et al., 2013; Supplemental Table S1), suggesting that they may be involved in the plant immune system. Therefore, to begin testing the role of these genes in PTI, a CRISPR mutagenesis screen targeting 54 of the 55 tomato LRR-XII genes was designed (Fig. 1). One family member, Solyc09g064950, was omitted because it is likely incorrectly annotated at only 192 bp long.

Three gRNAs were selected to specifically target each of the LRR-XII gene family members as well as a control phytoene desaturase (PDS), as a knockout would result in obvious photobleaching (Liu et al., 2002; Supplemental Table S1). The 165 gRNAs were synthesized, pooled, and cloned into a Cas9 binary vector using DNA assembly (Jacobs and Martin, 2016). Sequencing of 105 Escherichia coli clones indicated that there was no obvious gRNA bias from the library preparation (Supplemental Table S2). Eight of the sequenced clones contained errors: four were considered critical, since the protospacer was truncated (three) or the Pol III poly-T termination signal was deleted (one), and four contained noncritical errors in the U6 promoter or scaffold that are unlikely to impair function. With no apparent bias and a low critical error rate (4%), the LRR-XII library was cloned at 25× coverage (∼4,200 clones), and the resulting plasmid library was transformed into A. tumefaciens LBA4404 en masse at high coverage (∼250×) for stable transformation of the easily transformed M82 tomato background.

Thirty-one T0 events were regenerated and transplanted in the greenhouse. (Multiple clones were obtained for several of the transgenic lines, but as all results were identical between clones, only results from a single clone are reported.) The gRNAs were PCR amplified and Sanger sequenced from each event. In total, 24 of the events contained 27 gRNAs (Table I), with 25 unique gRNAs (two gRNAs were observed twice). Multiple gRNAs were incorporated in two lines, with events LRR-4 and LRR-24 containing three and two gRNAs, respectively.

Table I.

Summary of mutant lines from the LRR-XII library

Mismatches between the spacer and M82 target sequences are underlined. Alleles are combined data from T0 and T1 plants. The genotype of the plants is indicated. Active indicates that Cas9 is actively producing new alleles in those lines. NA, Not applicable; WT, wild type.

Line Target + PAM Target LRR-XII Family? T0 and T1 Alleles Genotype Notes
LRR-1 GAAGATATGGTTACTCATGTTGG Solyc07g016010 Yes −4 bp deletion/WT, active M82sp+ T1: some weak plants, low seed set, sterility
LRR-2 GTCACCACAAAGTGCATCATTGG Solyc06g006040 Yes −4/+609 (from Solyc06g006020) M82sp+ T0 and T1: high seed weight, low fruit set
GTCACCACAAAGTGCATCATCGGa Solyc06g006020 Yes −9/larger deletion
GTCACCACAAAGTGCATCATCGGa Solyc06g006050 Yes −1/no amplification (276 upstream)
GTCACCACAAAGTGCATCATCGGa Solyc06g006070 Yes −9/−8 (315 upstream)
LRR-3 No gRNA N/A N/A N/A M82sp+
LRR-4 GGCAATCTCAGCTTCCTAACAGG Solyc10g085110 Yes WT M82sp+ T1: leaf development (1/4)
GAGATGCAACATCCACATCTGGG Solyc03g118330 Yes −1/−2,112 bp
GCAGGTGACCAAGTCCATAAGGG Solyc04g014900 Yes −1/no amplification (unresolved)
LRR-5 GAAAGATCAAAAATTTCAGCAGG Solyc07g018180 Yes WT, sequence different from the reference M82sp+
LRR-6 GGATTAGTGATTCCTTTCCGGGG Solyc06g006020 Yes −2/−3/−7 M82sp+ Mostly parthenocarpic fruit with small aborted seed; one T1 plant
GGATTAGTGATTCCTTTCCGAGGa SL2.50ch00:17387212 No −5/−6/WT
LRR-7 GAACTCTCATCATTGTCGGAGGG Solyc03g096190 Yes WT M82sp+
LRR-8 GCTTTGGAATATCTTCACCACGG Solyc03g006100 Yes −1/−5 M82 T1: serrated cotyledons
LRR-9 GTAGGGATAGTGTCAACTATGGG Solyc08g075620 Yes −7/−1/−3 M82sp+ Sterile
LRR-10 GCTTTAGACTATCTCCATAAAGG Solyc02g068880 Yes −2/WT/−∼330 M82sp+ Sterile
LRR-11 GTGTACAAAGGTGCATTGTTTGG Solyc04g014890 Yes +1 homozygous M82 T1: stunted, bleached plants (1/4)
GTGTACAAAGGTGCATTGTTTGGa Solyc04g014650 Yes −2/+1
GTGTACAAAGGCACATTGTTTGGb Solyc04g014900 Yes +1/WT
GTGTACAAAGGCACATTGTTTGGb Solyc04g015000 Yes WT
GTGTACAAAGGCACATTGTTTGGb SL2.50ch04 6829670 No Active
LRR-12 GAAAGCAACTTGCTTGGTAACGG Solyc02g068830 Yes WT M82sp+
LRR-13 GCAGGCAAAAAACCGTCCAAGGG Solyc02g068820 Yes WT M82sp+
LRR-14 GTAGGGATAGTGTCAACTATGGG Solyc08g075620 Yes WT M82 T1: no anthocyanin (1/4)
LRR-15 GGCTACAGCTTGTTAATAGTCGG Solyc06g005880 Yes −7/−87 M82sp+
LRR-16 GGCCTATTCCGAGTTCATTCGGG Solyc02g031790 Yes −4/−5 M82sp+
LRR-17 GGTCAGATACCTAGTACAATCGG Solyc08g075600 Yes −1/−8 M82
LRR-18 GTCATTTTCATCCTTTCGACCGG Solyc02g072400 Yes WT M82
LRR-19 GGCGGCCAACCATCCACACTTGG Solyc10g085120 Yes −1/−3/−4/−5 M82sp+ Likely tetraploid
LRR-20 GAGATTTGAATTTGAGAAGCTGG Solyc02g072380 Yes WT M82sp+
LRR-21 GAGATATTAAACAGTGACGCGGG Solyc04g015980 Yes +1/−1/−4 M82sp+ Mostly parthenocarpic fruit with small aborted seed; one T1 plant
LRR-22 GCAGCAGCTTGTTATGACTTTGG Solyc06g006030 Yes WT M82sp+
LRR-23 GGTGAAATTCCTGCCGAGCTGGG Solyc02g070890 Yes −2 homozygous M82sp+ T1: all are insensitive to flg22 (expected)
GGTGAAATTCCTGCCGAACTAGGb Solyc02g070910 Yes Somatic active
LRR-24 GCAGGCAAAAAACCGTCCAAGGG Solyc02g068820 Yes WT M82sp+ T1: some plants light bleaching, slow-growing plants, low fruit set
GTGAGGTTCCCAATAGAAGGCGG Solyc07g018190 Yes WT
LRR-25 GTGCAGTTAGATAAACCAGGGGG Solyc02g072440 Yes −3/−4/−9/−10 M82sp+ T0 would not self, many parthenocarpic fruit, few seeds from LRR-25 × M82sp+
aUnintended targets that were missed during gRNA design.
bOff-targets to the gRNA sequence.

Mutations were detected in 15 of the 24 (62.5%) gRNA-containing T0 events and subsequent T1 lines, for a total of 24 mutated loci. Eighteen of the LRR-XII genes were mutated, and two of the genes were mutated twice (Solyc04g014900 and Solyc06g006020). There were also mutations outside of the LRR-XII gene family: two mutations were within ∼300 bp of LRR-XII family members, and two mutations were in intergenic regions. The types of mutations observed were standard for Cas9-induced nonhomologous end joining: short insertions and/or deletions were predominant, although larger deletions (LRR-4 and -10) and rearrangements (LRR-2) also occurred. All but three events appeared to have biallelic mutations (LRR-1, -10, and -11). The M82sp+ genomic sequence of the Solyc07g018180 target in the event LRR-5 contained several mismatches compared with the reference genome (Heinz 1706), which likely inhibited cleavage by Cas9 (Table I).

Heritability of On- and Off-Target LRR-XII Mutations

Seed was obtained from 22 of the T0 lines, and the alleles identified in T0 were observed segregating in T1. The results are summarized in Table I. New alleles were identified in T1 progeny derived from three lines (LRR-1, -11, and -23), indicating continued activity of the Cas9 nuclease. All other lines contained the same mutant alleles identified in T0, indicating that all wild-type alleles were mutated in the T0 generation and did not undergo additional mutations. Null segregants for the Cas9 T-DNA also were identified in the T1 generation. These transgene-free progeny contained the same alleles as the T0 parents and T1 siblings, further confirming the presence of heritable mutations.

Lines without mutations in T0 were checked for new mutations in T1, but no new on-target mutants were identified. However, two lines (LRR-11 and -23) contained off-target mutations in T1 that were absent in the T0 generation (Table I). LRR-11 SL2.50ch04 6829670 and LRR-23 Solyc02g070910 (Fls2.2; Rosli et al., 2013) had overlapping chromatograms only in T1 plants containing the T-DNA, indicating that continued activity of Cas9 resulted in the additional mutations. The LRR-23 Solyc02g070890 (Fls2.1; Rosli et al., 2013) gRNA has a single G-to-A mismatch at the corresponding Fls2.2 locus, 3 bp 5′ from the PAM site. No Fls2.2 mutations were recovered in Cas9-free T2 plants, even though their Cas9-positive T1 parent contained mutations, suggesting that the observed off-target mutations at Fls2.2 are somatic events.

A single mutant allele was detected in the T0 generation at five of the target loci, which suggested instant homozygosity at these targets. However, three of these putative homozygous loci could not be amplified by PCR from a subset of T1 individuals. By designing primers to amplify an ∼3-kb portion of the genomic region surrounding Solyc03g118330, a 2,112-bp deletion was detected in the LRR-4 T0 and T1 plants. In LRR-2 at the Solyc06g006040 target, a 609-bp insertion was observed only in T1 plants homozygous for the insertion, as the initial PCR conditions favored the smaller (4-bp deletion) product in heterozygous individuals. Several LRR-4 T1 individuals also failed to amplify Solyc04g014900, suggesting another allele that could not be observed. Despite numerous attempts to amplify around the targeted region, this allele could not be resolved. Therefore, true instant homozygosity appears to have occurred only at two of the targeted loci.

Unintended Targets

During mutation analysis, it became clear that the scoring parameters used to select the LRR-XII gRNAs incorrectly scored two of the incorporated gRNAs, leading to perfectly matched, unintended targets (LRR-11 Solyc04g014650 and LRR-6 SL2.50ch00:17387212). Both unintended targets had heritable mutations, one of which is in the LRR-XII gene family (Table I).

Errors in the reference genome also contributed to unintended targets. There are multiple contig breaks in the tomato reference genome near and within several of the LRR-XII gene family members. While the LRR-2 and LRR-6 gRNAs were specific to the reference sequence, perfect target sites are in fact present in the tomato genome. For the LRR-2 gRNA, sequence data from a BAC contig (AC254430) revealed three perfect target sites, one in Solyc06g006020 and two that are ∼300 bp upstream of the predicted start codons of Solyc06g006050 and Solyc06g006070 (Supplemental Fig. S1). Sequence analysis of each of these sites in the LRR-2 line indicated that heritable mutations were present at each of these unintended targets. Interestingly, the 609-bp insertion at the Solyc06g006040 target site is an inversion of part of the first intron of Solyc06g006020. The exact alleles produced at all the target sites have not been completely resolved. As these four genes occur on a single cluster on chromosome 6, it is likely the simultaneous DSBs resulted in larger chromosomal alterations that prevent PCR amplification. LRR-6 also contains a mutation in a locus adjacent to the target gene Solyc06g006020 (Table I). Given the similarity between the two sequences, it is possible that this locus was produced by a tandem duplication. However, it is unclear if this locus represents an unidentified LRR-XII family member or is a pseudogene.

Fls2.1 Is Likely Required for Age-Dependent PTI in Tomato

There are two tandemly duplicated homologous genes in tomato of the Arabidopsis flg22 receptor gene FLS2. The line LRR-23 contains a 2-bp deletion in one of these genes, Fls2.1, resulting in a premature stop codon in the third LRR domain. A single 2-bp deletion was observed in T0 as well as all 19 T1 plants sequenced, indicating a true homozygous event. As described above, somatic mutations were observed in the off-target site on Fls2.2 in T1 plants containing Cas9. All 29 T1 plants tested failed to respond to flg22 in a reactive oxygen species (ROS) assay, while the response to flgII-28, which is conferred by Fls3, was maintained (Fig. 2A; Supplemental Fig. S2A). When vacuum inoculated with Pseudomonas syringae DC3000∆avrPto∆avrPtoB, there were greater disease symptoms (Fig. 2B) and an approximately 1-log increase in bacterial growth in T1 LRR-23 plants compared with the M82sp+ wild-type control (Fig. 2C; Supplemental Fig. S2, B and C). The difference in bacterial growth was greatest, but more variable, in younger leaves, as has also been observed in Arabidopsis fls2 mutants (Zipfel et al., 2004). Together, these data indicate that Fls2.1 plays the primary role in the response to flg22 in tomato leaves.

Fls2.1 is required for flg22 perception and disease resistance. A, Alignment of the wild-type (wt) Fls2.1 target sequence and a 2-bp deletion in LRR-23. The protospacer is in blue, and the PAM is in red. B, Leaf discs were treated with the peptide flg22 or flgII-28, and the production of ROS was measured over time. Values are averages ± sd of 14 LRR-23 T1 plants and four M82sp+ plants. Four leaf discs per plant were used for flg22, and two leaf discs per plant were used for flgII-28. Graphs were generated by GraphPad Prism 7. RLU, Relative light units. C, Disease symptoms on LRR-23 T1 plants 3 d after vacuum infiltration with 3.5 × 105 colony-forming units (CFU) mL−1 P. syringae DC3000∆avrPto∆avrPtoB. D, Bacterial growth measured on leaves at different developmental stages. Leaf discs were taken 2 d after vacuum infiltration with 3.5 × 104 CFU mL−1 DC3000∆avrPto∆avrPtoB. Bars are averages ± sd for six plants per genotype. Individual values are represented as dots. Graphs were generated by GraphPad Prism 7.

Aberrant Phenotypes in the LRR-XII Lines

Six T1 lines had obvious segregating aberrant phenotypes. A subset of T1 progeny from LRR-1 had thin, spindly plants with long internodes and failed to produce flowers (Supplemental Fig. S3, A and B), plants from LRR-4 terminated after the fourth true leaf with wiry leaflets (Supplemental Fig. S3, C and D), LRR-8 had serrated cotyledons (Supplemental Fig. S3E), LRR-11 had pale seedlings that usually died a few weeks after germination or remained dwarfed (Supplemental Fig. S3F), and LRR-14 had anthocyanin-free plants (Supplemental Fig. S3G). Segregation analysis suggested that these characteristics were simple recessive traits (data not shown). However, none of the observed mutations cosegregated with these phenotypes (Supplemental Fig. S3H). Off-target mutations were checked in lines LRR-1, -4, -8, -11, and -14. Line LRR-14 did not contain any mutations at the target locus in the T0 or T1 generation. While line LRR-11 did contain off-target mutations (Table I), none of the mutations segregated with the pale/dwarf phenotype. Because of the stunted, bleached phenotype in this line, mutations were checked at each of the three PDS target sites included in the gRNA library, and no mutations were found. The presence/absence of the T-DNA also was evaluated, but again, there was no association of the T-DNA insertion with the aberrant phenotypes (Supplemental Fig. S3H).

We considered the possibility that some of the aberrant phenotypes in the LRR lines could be explained by transient T-DNA mutations (i.e. transient expression of the T-DNA, without genomic integration), as we observed in the transporter lines (see below). As transient mutations could occur at least at any of the 165 target loci, spread across 55 genes that are approximately 3 kb in length, we opted to use the long read length of single molecule, real time (SMRT) sequencing to screen for these mutations. Fifty-eight amplicons covering the 54 LRR-XII genes and four potential off-target genes were sequenced from T0 and T1 plants in LRR-1, -4, -11, and -14. The T1 plants selected for sequencing showed the respective aberrant phenotypes and, therefore, should be homozygous for the mutated genes. Across all samples, at least five aligning circular consensus sequences (CCS) were obtained for 42 to 55 of the 58 amplicons (Supplemental Table S3). All but one of the previously identified mutations were detected, indicating that this custom amplicon-sequencing approach was suitable for identifying mutations (Supplemental Table S4). While we did observe some candidate mutations at gRNA target sites, none were homozygous in the T1 individuals and are likely false positives. Based on these observations, it seems most likely that the aberrant phenotypes are due to somaclonal variation resulting from the transformation and tissue culture processes (Larkin and Scowcroft, 1981; van den Bulk et al., 1990).

Higher Order Library

Having demonstrated that transformation with a pooled CRISPR library could generate a targeted collection of mutant lines, a second CRISPR library was designed. To increase throughput, plasmids containing three gRNAs were constructed. The library also was limited to 18 genes, to increase the probability of obtaining multiple independent mutations in each of the target genes, and with two gRNAs per gene to reduce the cost of oligonucleotide synthesis. The 18 targeted genes encode putative transporters, and we reasoned that some of these may be susceptibility factors, as their expression is suppressed by the plant during PTI (perception of flgII-28; Rosli et al., 2013) and induced by the pathogen effectors AvrPto and AvrPtoB, potentially to promote access to water and/or nutrients.

For library construction, the 36 gRNAs were divided into three pools with six genes targeted per pool. Three separate DNA assembly reactions were used to generate the pools, and the reactions were used as templates for PCR with gRNA-position-specific primers (Fig. 3A). During initial construction attempts, it became apparent that the PCR amplification step was biasing the relative abundance of the gRNAs. Therefore, a lower number of PCR cycles (14) was used to minimize this bias. The final assembly was cloned into E. coli at 118× coverage (∼1,400 clones), introduced into A. tumefaciens, and transformed into the tomato genotype Rio Grande-PtoR.

Higher order transporter library scheme. A, Twelve gRNAs, targeting six genes, were pooled and assembled per gRNA position. Amplification with primers containing unique nucleotide sequences (Torella et al., 2014) allowed for specific positioning of the gRNAs in the final DNA assembly step. B, All 59 transporter events were checked for mutations at all 36 targeted loci and incorporation of the T-DNA. Average mutant values are reported for integrated gRNAs (Int. ∆), total transient mutations (Tran. ∆), and total mutations (∑∆).

Numerous Combinations of Mutated Genes Were Obtained

Fifty-nine unique events (transporters 1–59) were obtained with the transporter library (Fig. 3B). Three events did not contain mutations or the T-DNA and are likely nontransformed escapes. Thirty-five of the 36 gRNAs were observed in the T0 plants (Supplemental Table S5). The distribution of the gRNAs was much greater than in the LRR-XII library, likely due to the PCR amplification step during library construction. For every event, all of the 36 targeted loci were amplified and Sanger sequenced. Forty-two events (71%) contained mutations, with each event containing a unique set of mutant-gene combinations. Fifteen out of the 18 transporter genes were mutated at least once, and 14 of the genes were mutated at least three times (Fig. 3B; Supplemental Table S5). Seven of the events contained T-DNAs with single gRNAs, with each of these coming from the third gRNA position and none resulting in mutations. Ten (4%) of the incorporated gRNAs had errors in the protospacer that likely prevented mutations at their target sites.

There was a wide range of editing efficiencies for the different gRNAs. Sixteen of the incorporated gRNAs had no mutations at all, although seven of these were only counted two or fewer times, and their failure to induce mutations may be attributed to a small sample size. Approximately half of the total mutations were derived from gRNAs in the first position, 31% from the second, and 17% from the third (Supplemental Table S5), suggesting that there may be a position effect within the gRNA array. Two genes that were not mutated, Solyc06g060620 and Solyc08g066940, had 11 and 16 gRNAs, respectively, suggesting that these are ineffective gRNAs. Alternatively, there might be selection against mutations in these genes. Thus far, 21 T1 lines have been checked, and heritable mutations have been confirmed in 19 of them (Supplemental Table S6).

Transiently Expressed T-DNA-Induced Mutations

Ten of the 59 events contained mutations at gRNA target sites in which the respective gRNAs were not observed after sequencing the incorporated gRNAs (Fig. 3B). All mutations were located between the third and fourth nucleotides upstream of the PAM, a hallmark of Cas9-mediated gene editing. These results suggest that transient expression of the T-DNA, without genomic integration, led to these mutations. Such possible transient T-DNA mutations accounted for 23 of the 93 mutations (25%; Supplemental Table S5). For example, no gRNAs could be amplified from transporter-5, -10, and -11, but mutations were detected in each of them and in the T1 progeny of transporter-5 and -11. Since rearrangements of the incorporated T-DNA might prevent amplification of the gRNA amplicons used for sequencing, and therefore, detection of gRNAs, PCR with primers designed to amplify a minimal portion of the gRNA (227 bp) was performed. Again, these three events tested negative (data not shown). These events also lack the two Cas9 amplicons, and transporter-10 and -11 failed to amplify the nptII selectable marker, suggesting that these events are escapes. The seven other transient T-DNA events contained other gRNAs, with and without mutations at their target genes.

Independent Events with Reproducible Phenotypes

Two different aberrant phenotypes were observed in independent T0 transporter events. Transporter-3 and -29 have biallelic mutations in Solyc06g071500 (Fig. 4C), a homolog of the Arabidopsis boron transporters BOR1 and BOR2 (Takano et al., 2002; Miwa et al., 2013). Both T0 plants had developmental phenotypes reminiscent of tomato plants with boron deficiency (Uraguchi et al., 2014), namely leaf curling, leaves that fail to expand, and flowers that fail to develop (Fig. 4A). Transporter-8 has a monoallelic mutation at Solyc06g071500, leaving one wild-type copy intact, and symptoms were milder and largely restricted to lateral shoots. Transporter-17 has no apparent phenotype, consistent with only ∼10% of its alleles being modified, indicating that this is a chimeric event (Fig. 4C). Daily spraying of transporter-3 and -29 plants with 100 µm boric acid rescued the phenotype (Fig. 4B) and allowed the plants to flower and set fruit. T1 plants from transporter-3 all showed the same boron deficiency symptoms and had either 31- or 4-bp deletions (Supplemental Table S6). T1 plants from the heterozygous transporter-8 segregated for the boron deficiency phenotype (Fig. 4D). Nine T1 plants had the phenotype and also showed an assortment of frame-shift alleles not observed in the T0 plant (Fig. 4E). One normal T1 plant had a homozygous, in-frame 3-bp deletion. Two of the normal plants shared the same heterozygous genotype as the T0.

Mutations in Solyc06g071500 are associated with a boron-deficient phenotype. A, The T0 transporter-3 showed a strong aberrant phenotype similar to boron deficiency. B, Daily foliar treatments with 100 µm boric acid allowed the plant to recover and set fruit. C, Events with boron-deficient phenotypes are associated with mutations in Solyc06g071500. Estimated allele frequencies are from TIDE analysis. D, Transporter-8 T1 plants segregate for the boron-deficient phenotype. E, Normal T1 plants contain a functional wild-type (WT) copy or an in-frame deletion, and all boron-deficient plants contain only frame-shift alleles.

A second aberrant phenotype was observed in the transporter-10 and -38 plants. Both of these plants had more open architectures and wavy or rippled leaves and were sterile (Fig. 5A). These events share monoallelic mutations in Solyc08g082990 and biallelic mutations in Solyc05g005940 and Solyc04g005070 (Fig. 5C). Solyc08g082990 is most likely not the causative gene, as both events have monoallelic mutations. There are 11 other events with mutations in Solyc05g005940, nine of which contain biallelic mutations (Fig. 3B; Supplemental Table S6), and none showed the same phenotype, ruling out this gene. There are only two other events (transporter-2 and -26) with mutations in Solyc04g005070. Transporter-2 is monoallelic, and transporter-26 has an in-frame, 6-bp deletion. The heterozygous condition in these events likely explains the lack of phenotype in these plants. A transporter-2 T1 plant with a homozygous 1-bp deletion showed the same wrinkled leaf phenotype (Fig. 5), further reinforcing the association of Solyc04g005070 with the aberrant phenotype.

Mutations in Solyc04g005070 are associated with an aberrant phenotype. A, T0 transporter-38 showing the open architecture, elongated internodes, and rippled or wavy leaves. B, Transporter-10 and -38 have the same phenotype and are the only events with biallelic mutations in Solyc04g005070. WT, Wild type. C, A transporter-2 T1 progeny with a homozygous 2-bp deletion shows the same aberrant phenotype.

DISCUSSION

Mutant Collections Are Efficiently Generated by Pooled A. tumefaciens Transformation

The ability to generate knockout mutations with CRISPR is relatively straightforward in plant species competent for transformation and regeneration (Bortesi and Fischer, 2015). Most attempts usually focus on generating several independent events with mutations in one or a few genes at a time. However, this approach can be wasteful, as there are often more mutant lines generated than are needed. By pooling large numbers of CRISPR plasmids for transformation of tomato, we efficiently obtained 15 heritable mutations out of the 54 LRR-XII genes targeted from a single transformation experiment. By increasing the gRNA payload of each T-DNA and decreasing the number of gene targets, it was possible to obtain multiple mutations in all but three of the 18 targeted transporter-encoding genes. Importantly, 14 of the 18 targeted genes were mutated at least three times. This high level of replication allowed the association of two of the putative transporter genes with two aberrant phenotypes at the T0 and T1 generations. The replication also will be valuable to confirm associations as this population is screened in the future.

The use of pooled CRISPR libraries for plant transformation presents an alternative method to standard genetic screens that rely on EMS, fast neutrons, or transposon tagging. These methods generally require screening thousands of M2 families to identify lines with the desired phenotype. Hybridization and genetic mapping is then used to obtain a list of candidate genes that can be followed up with reverse genetic approaches. By comparison, a pooled CRISPR library can generate mutations in a specific set of genes, targeting a gene family or those identified in an RNA-seq data set. Ideally, a T0 population of sufficient size would be generated that is then used to make T1 lines for the screen, although some phenotypes are apparent as early as T0. Approximately 60% to 70% of the T0 plants likely contain the desired mutations, and the T1 plants can be screened for the desired phenotype. Candidate lines can then be rapidly characterized by sequencing the incorporated gRNAs and/or the predicted target sites. This approach could be used to complement standard EMS-based screens or even as a follow-up to quickly narrow down a list of candidate genes from a screen or mapping experiment.

The various lines derived from the LRR-XII library demonstrated the utility of obtaining a relatively large number of knockout lines with a single transformation experiment. The presence of apparent somaclonal mutants hindered the evaluation of these lines, but such effects can be overcome by backcrossing to the parental line. After the generation of the LRR-XII mutant library, it was clear that there was an advantage of having multiple targets within a single event to generate additional mutants. This led to the design of a higher order library and a smaller number of gene targets to allow for the isolation of multiple mutants per gene target, which also can ameliorate the issue of somaclonal variation.

In contrast to the single FLS2 gene in Arabidopsis, tomato has two tandemly duplicated genes, Fls2.1 and Fls2.2. However, mutations in the Fls2.1 or Fls2.2 genes have not been reported previously, and it was unknown if they have a redundant function in flg22 recognition. We found that a mutation in just Fls2.1 was sufficient to knock out the response to flg22 in a ROS assay and to compromise host resistance against P. syringae. A noteworthy result was the correlation between leaf age and bacterial growth under our experimental conditions, where the younger leaves were most susceptible to the loss of Fls2.1 and the older leaves showed only slight differences (Fig. 2D; Supplemental Fig. S1, B and C). Visual inspection of the disease symptoms in the LRR-23 plants was consistent with the bacterial growth difference (data not shown). This is despite the fact that the perception and response of flg22 are not functional in the older, fully expanded leaves of LRR-23 T1 plants, as determined by a ROS assay (Fig. 2B). Consistent with previous observations in Arabidopsis (Zipfel et al., 2004), these results suggest that it will be important to consider the developmental stage of the mutant plants when screening for possible differential responses to pathogens or peptide elicitors.

The transporter collection of mutants also can be screened with different pathogens to test the hypothesis that these genes may be susceptibility factors. The ability to make multiplex mutations in genes with potential functional redundancy may provide a unique advantage to this type of screen over other mutagenesis systems where there are few options to deal with genetic redundancy.

Transient Expression

The use of pooled A. tumefaciens strains permitted the identification of genes that were mutated during apparent transient expression of the T-DNAs. Other groups have reported successful targeted mutations without the incorporation of DNA or RNA with biolistic delivery methods or with biolistic or protoplast transfection of ribonucleoprotein complexes (Woo et al., 2015; Svitashev et al., 2016; Zhang et al., 2016; Liang et al., 2017). The observations of transiently expressed T-DNA leading to CRE recombinase activity (Ghedira et al., 2013) and Cas9-induced DNA mutations reported here and by others (Iaffaldano et al., 2016) indicate that A. tumefaciens-mediated delivery of T-DNA is another viable means of obtaining transgene-free gene-edited plants. This may be useful for plant species that are better suited to A. tumefaciens-mediated transformation and regeneration over biolistic or protoplast approaches. It will be interesting to determine the rate of mutagenesis in the absence of selection or if selection against events with incorporated T-DNAs will be needed. Alternatively, lines with a knockout in polymerase Θ could be utilized to produce plants resistant to stable integration but still support transient T-DNA expression of nucleases (van Kregten et al., 2016).

However, the presence of transient T-DNA modifications does present a limitation to the pooled library approach. A library with a large number of gene targets may be desirable under certain conditions (e.g. species, phenotype, transformation efficiency, etc.), but any candidate lines would need to be checked at every target gene because sequencing the incorporated gRNAs would not identify all possible mutation sites. In the transporter library, it would not have been possible to so rapidly associate Solyc04g005070 with the aberrant phenotype observed in transporter-10 and -38 if only the incorporated gRNAs were considered. Such transient T-DNA mutations may not be a problem in all A. tumefaciens-mediated plant transformation systems where transient effects are limited (Ghedira et al., 2013).

Practical Considerations

Genome-wide CRISPR libraries are routinely used in animal cell culture systems (Shalem et al., 2015). While it is tempting to extend the pooled A. tumefaciens approach presented here to a genome-wide screen in plants, there are a number of technical and practical limitations that would need to be overcome: namely, the occurrence of transient T-DNA mutations. If every potential gRNA target site needs to be checked in every plant, a significant amount of time and money would be required for each plant. However, if transient T-DNA modifications could be eliminated, or reduced, only targets identified by the incorporated gRNA(s) would need to be checked. This issue also may be overcome by using a sufficiently large population or if sequencing costs are minimal.

Identifying incorporated gRNAs becomes a laborious task when multiple T-DNAs are incorporated into an event or the same Pol III promoter is repeated, as was done for the transporter library. Several rounds of cloning and Sanger sequencing were required, and we were still unable to confidently identify all gRNAs in two of the T0 plants (Supplemental Table S7). A gRNA amplicon-sequencing approach would be more practical for large numbers of events.

While CRISPR systems are generally quite efficient in plants, the results from the transporter library clearly show that there is room for improvement in gRNA and/or vector design. Optimization of library preparation by further reducing the number of PCR amplification cycles for a more even gRNA distribution would likely have improved the outcome. Certainly, the use of more efficient gRNAs also would be an advantage: 13 of the gRNAs resulted in no mutations at all, and many of the most abundant gRNAs had mutation rates lower than 20%. However, despite the variety of gRNA scoring algorithms available for animal systems (Haeussler et al., 2016), there are currently no plant-validated gRNA activity predictions available.

There also is an apparent negative correlation between gRNA position and gRNA efficiency. In position 1, the gRNAs had an average efficiency of 40%, in position 2, 32%, and in position 3, 14%. This trend also is consistent for transient T-DNA mutations: 45% of all transient mutations came from position 1, 36% from position 2, and 18% from position 3. In all, 51% of all mutations came from the first position (Supplemental Table S5). Multiplexing with CRISPR arrays using identical Pol III promoters is a common strategy (Belhaj et al., 2013; Endo et al., 2016; Vazquez-Vilar et al., 2016), but none have reported a decrease in efficiency associated with gRNA position in the array. It is not immediately clear why the repeated use of a Pol III promoter would lead to a decrease in efficiency across the array. The gRNA array appears to be stably integrated, as all but two of the events (transporter-13 and-27) had the same number of gRNAs in each pool. If gene silencing were the cause, then the effects should be spread across all gRNA cassettes. The use of different Pol III promoters in a gRNA array might improve the mutation efficiency and would be a benefit when amplifying the gRNAs from an array, as a PCR across the entire array results in a ladder of products and was part of the reason why identifying the gRNAs was challenging. It will be worthwhile to test a number of hypotheses from this observation, as the use of multiplexing CRISPR reagents will surely expand in the future.

The quality of the reference genome is an important consideration when setting up a CRISPR screen. In plants, LRR genes are well known to occur in clusters of tandemly duplicated genes, which can make their assembly and annotation difficult (Andolfo et al., 2014). We experienced this issue as the guide in LRR-2 has (at least) four targets in a single cluster on chromosome 6. These unintentional targets would have been overlooked if not for the insertion of the Solyc06g006020 intron into Solyc06g006040.

Off-Targets Increase the Throughput of a CRISPR Screen

Most discussions of off-targeting in the context of gene editing focus on limiting off-target effects. However, as we have observed and has been shown in rice (Oryza sativa; Endo et al., 2015), off-targets can be a benefit in some cases by increasing the multiplexing capacity of the screen. In the LRR-XII library, nonspecific gRNAs allowed the recovery of six additional heritable mutations, three of which were in the LRR-XII genes. In a screen, specificity is less important than the number of genes that can be screened simultaneously. If multiplexed events show a particularly interesting phenotype, it is straightforward to segregate unlinked mutations by backcrossing or to retransform with gene-specific gRNAs. Future library screens can be easily designed with nonspecific gRNAs in mind.

CONCLUSION

The generation of knockout mutant plant lines is important for functional genomic studies. The high efficiency of mutagenesis of CRISPR systems in plants permits the rapid development of knockout plant lines. Stable transformation methods usually target one gene at a time and lead to the development of more mutant lines than necessary. The approach reported here generates many mutagenized lines from relatively few transformation attempts. Various combinations of mutants also can be obtained in cases of gene redundancy and may be useful for various screening strategies. Importantly, this approach can be applied to any plant transformation system.

MATERIALS AND METHODS

Designing gRNA Targets and Library Cloning

The commercial software Geneious R8 (Biomatters; http://www.geneious.com/; Kearse et al., 2012) was used to identify gRNAs for each of the 54 LRR-XII gene family members using the target motif GN20GG. The tomato (Solanum lycopersicum) genome reference sequence (SL2.50) was used as an off-target database using the score all sites option. For each gene, the target sequences and their coordinates were extracted, and a custom Perl script was used to select the top three nonoverlapping gRNAs for each gene. Oligonucleotides were designed with the GN19 sequences plus 20-nucleotide overlapping regions of the MtU6 promoter and scaffold for DNA assembly with NEBuilder HiFi DNA Assembly Master Mix (New England Biolabs), as described previously (Jacobs and Martin, 2016). Oligonucleotides were synthesized and pooled in equimolar amounts by Integrated DNA Technologies.

The DNA assembly mix was transformed into electrocompetent Escherichia coli 10β cells (New England Biolabs). Two DNA assembly reactions were used for 10 transformations to obtain approximately 4,000 clones representing 25× coverage of the library. A total of 114 individual clones were picked for colony PCR to estimate the distribution of the library. All of the ∼4,000 colonies were scraped from the plates and used for plasmid extraction using EconoSpin Mini spin columns (Epoch Life Sciences). A diagnostic digest using the restriction enzymes StyI-HF and EcoRV-HF (New England Biolabs) was performed as described previously (Jacobs and Martin, 2016) on the pooled plasmids to ensure that there were no obvious cloning errors. The plasmid pool was transformed into electrocompetent Agrobacterium tumefaciens LBA4404 cells. The resulting transformation was plated at 250× coverage, scraped together and resuspended in Luria-Bertani medium, and saved as a glycerol stock.

To construct the three-gRNA vectors, two DNA assembly reactions were performed as described previously (Jacobs and Martin, 2016). The first was between the oligonucleotide pools, MtU6 promoter, and scaffold (Fig. 3). There were three reactions for the three separate positions (pools). The reactions were then used as templates for PCR using the MtU6 and scaffold primers with overlapping unique nucleotide sequences. PCR products were purified with DNA Clean & Concentrator-5 columns (Zymo Research) and used in a second DNA assembly. The DNA assembly mix was transformed into electrocompetent E. coli 10β cells, and ∼1,400 colonies were recovered (118× coverage). Sequencing the three gRNAs from 18 clones ensured that there was no obvious gRNA bias. This aspect was crucial, as initial attempts to clone the library had a clear overrepresentation of certain gRNA sequences. Reducing the number of PCR cycles to 14 resulted in a more even coverage of gRNAs.

Tomato Genotypes and Transformation

Tomato transformation was performed at the Boyce Thompson Institute transformation facility as described previously (Frary and Van Eck, 2005). The seed used for transformation was a mixture of M82 (determinate) and M82sp+ (indeterminate; Krieger et al., 2010). The primers sp175F and sp1088R (Supplemental Table S8) were used to amplify Solyc06g074350, and lines with sequences identical to the reference were scored as M82 and those differing were scored as M82sp+ (Table I). The genotypes were confirmed by growth habit in the greenhouse. The transporter library was transformed into the tomato genotype Rio Grande-PtoR by the plant transformation facility at the University of Nebraska. The E. coli library was used in a triparental mating to obtain a library of A. tumefaciens. Rescued plasmids from 10 clones were sequenced to ensure that no plasmid was overrepresented.

Genotyping and Mutation Analysis

DNA was extracted from tomato leaves using a modified CTAB method (Murray and Thompson, 1980; 20 g L−1 CTAB, 100 mm Tris, 81.82 g L−1 NaCl, and 2 mm EDTA, pH 6). Primer sequences can be found in Supplemental Table S8. PCR was performed with either GoTaq master mix (Promega) or Apex taq red Master Mix (Genesee Scientific) in 10-µL reactions using the recommended PCR conditions. Sanger sequencing was performed at the Cornell Biotechnology Resource Center. Chromatograms were aligned to target sequences in Geneious. Overlapping chromatograms at the cleavage site indicated a potential mutation, and the online program TIDE was used to estimate mutations and allele frequencies (Brinkman et al., 2014).

Oxidative Burst Assays

ROS assays were performed as described previously (Clarke et al., 2013) with slight modifications. Briefly, 100 nm of each peptide was used to elicit the ROS response. Four leaf discs per plant were used for flg22 treatment, and two leaf discs per plant were used for water and flgII-28 controls.

Plant Inoculations

Bacterial growth and disease assays were performed as described previously (Mathieu et al., 2014) with slight modifications. Pseudomonas syringae pv tomato DC3000∆avrPto∆avrPtoB was grown for ∼24 h on Kings B solid medium supplemented with rifampicin. Bacteria were scraped from the plate and resuspended in sterile 10 mm MgCl2 and diluted to OD600 = 0.35. Three milliliters of this suspension was used to inoculate 3 L of 10 mm MgCl2 containing 0.002% Silwet L-77, resulting in a solution of bacteria at 3.5 × 105 CFU mL−1. Plants were submerged in the bacterial solution, a vacuum of −80 kPa was applied for 2 min, and plants were allowed to dry standing up before being transferred to a growth chamber. Visual inspection of disease and photographs were taken 3 d after infiltration. For bacterial growth assays, three independent experiments were performed over 1 month. For each experiment, six plants per genotype of approximately equal size were used. A lower inoculum of 3.5 × 104 CFU mL−1 was used, and 2 d after infiltration, three leaf discs were taken from each leaf with a 6-mm-diameter cork borer for the youngest leaves and an 8-mm-diameter cork borer for the other leaves. Leaf samples were ground in 1 mL of sterile 10 mm MgCl2, and serial dilutions were plated to count bacterial colonies.

Boric Acid Treatments

A solution of 100 µm boric acid (Sigma-Aldrich) was sprayed daily on T0 events transporter-3 and -29 such that the leaves were completely wet with the solution.

SMRT Sequencing

Primers for the amplicons used for SMRT sequencing can be found in Supplemental Table S8. Amplicons were designed to be 3 to 5 kb in length and validated by Sanger sequencing and alignment to the reference sequence. Each gene-specific forward and reverse primer was appended with a 5′ block and the PacBio Universal Sequences (forward primer, 5′-GCAGTCGAACATGTAGCTGACTCAGGTCAC-3′; and reverse primer, 5′-TGGATCACTTGTGCAAGCATCACATCGTAG-3′). Approximately 10 ng of genomic DNA was used as a template for PCR and amplified in 10-µL reactions with 2x HotStart Ready Mix HiFi polymerase (KAPA Biosystems) and 0.2 µm of each primer under the following conditions: 95°C for 3 min; 32 cycles of 98°C for 20 s, 62°C for 30 s, and 72°C for 4 min; followed by 72°C for 5 min. Each amplicon was amplified separately and confirmed by electrophoresis. Within each line, approximately equal amounts of amplicons were pooled, and products were column purified. A second PCR was used to add indexing primers (PBbarcode; Supplemental Table S8) with universal sequences (Full-Length 16S Amplification, SMRTbell Library Preparation and Sequencing; Pacific Biosciences) with 200 pg of purified amplicons under the following conditions: 20-μL reactions at 95°C for 3 min; 15 cycles of 98°C for 20 s, 62°C for 30 s, and 72°C for 4 min; followed by 72°C for 5 min. Amplification was verified by agarose gel and column purified. All amplicons were then pooled in equal concentrations and sent to the Icahn School of Medicine at Mount Sinai for sequencing. After sequencing, reads were demultiplexed, and approximately 1,000 CCS reads were generated per line. The reads were aligned to the 58 target amplicons using the Geneious function Map to Reference using the default conditions under the sensitivity setting High Sensitivity/Medium. Alignments were inspected manually for mutations at all target sites.

Supplemental Data

The following supplemental materials are available.

Acknowledgments

We thank Brian Bell, Jay Miller, Nick Vail, Jacob Wszolek, and Lesley Middleton for plant care; Laurens Pauwels for helpful comments on the article; Hernan Rosli for RNA-seq data analysis; Noe Fernandez-Pozo and Surya Saha for bioinformatics support; Joyce Van Eck and Patricia Keen for the LRR-XII library transformation; Thomas Clemente and Shirley Sato for the transporter library transformation; Robert Sebra, Diane Castillo, and Leah Newman for assistance with the design, sequencing, and analysis of the SMRT amplicons; Konrad Thorner and Samantha Mainiero for technical help; and Simon Schwizer, Christine Kraus, and Sarah Hind for technical advice.

Notes

Glossary

CRISPR clustered regularly interspaced short palindromic repeats
gRNA guide RNA
PAM protospacer adjacent motif
DSB double-stranded break
PTI pattern-triggered immunity
ROS reactive oxygen species
SMRT single molecule, real time
CCS circular consensus sequences
CFU colony-forming units

Footnotes

1This work was supported by the National Science Foundation (grant nos. IOS-1451754 and IOS-1546625 to G.B.M.).

References

  • Andolfo G, Jupe F, Witek K, Etherington GJ, Ercolano MR, Jones JD (2014) Defining the full tomato NB-LRR resistance gene repertoire using genomic and cDNA RenSeq. BMC Plant Biol 14: 120. [PMC free article] [PubMed] [Google Scholar]
  • Belhaj K, Chaparro-Garcia A, Kamoun S, Nekrasov V (2013) Plant genome editing made easy: targeted mutagenesis in model and crop plants using the CRISPR/Cas system. Plant Methods 9: 39. [PMC free article] [PubMed] [Google Scholar]
  • Berardini TZ, Reiser L, Li D, Mezheritsky Y, Muller R, Strait E, Huala E (2015) The Arabidopsis Information Resource: making and mining the “gold standard” annotated reference plant genome. Genesis 53: 474–485 [PMC free article] [PubMed] [Google Scholar]
  • Bortesi L, Fischer R (2015) The CRISPR/Cas9 system for plant genome editing and beyond. Biotechnol Adv 33: 41–52 [PubMed] [Google Scholar]
  • Brinkman EK, Chen T, Amendola M, van Steensel B (2014) Easy quantitative assessment of genome editing by sequence trace decomposition. Nucleic Acids Res 42: e168. [PMC free article] [PubMed] [Google Scholar]
  • Brooks C, Nekrasov V, Lippman ZB, Van Eck J (2014) Efficient gene editing in tomato in the first generation using the clustered regularly interspaced short palindromic repeats/CRISPR-associated9 system. Plant Physiol 166: 1292–1297 [PMC free article] [PubMed] [Google Scholar]
  • Burch-Smith TM, Anderson JC, Martin GB, Dinesh-Kumar SP (2004) Applications and advantages of virus-induced gene silencing for gene function studies in plants. Plant J 39: 734–746 [PubMed] [Google Scholar]
  • Carter JD, Pereira A, Dickerman AW, Veilleux RE (2013) An active ac/ds transposon system for activation tagging in tomato cultivar m82 using clonal propagation. Plant Physiol 162: 145–156 [PMC free article] [PubMed] [Google Scholar]
  • Char SN, Neelakandan AK, Nahampun H, Frame B, Main M, Spalding MH, Becraft PW, Meyers BC, Walbot V, Wang K, et al. (2017) An Agrobacterium-delivered CRISPR/Cas9 system for high-frequency targeted mutagenesis in maize. Plant Biotechnol J 15: 257–268 [PMC free article] [PubMed] [Google Scholar]
  • Clarke CR, Chinchilla D, Hind SR, Taguchi F, Miki R, Ichinose Y, Martin GB, Leman S, Felix G, Vinatzer BA (2013) Allelic variation in two distinct Pseudomonas syringae flagellin epitopes modulates the strength of plant immune responses but not bacterial motility. New Phytol 200: 847–860 [PMC free article] [PubMed] [Google Scholar]
  • De Buck S, Podevin N, Nolf J, Jacobs A, Depicker A (2009) The T-DNA integration pattern in Arabidopsis transformants is highly determined by the transformed target cell. Plant J 60: 134–145 [PubMed] [Google Scholar]
  • Depicker A, Herman L, Jacobs A, Schell J, Vanmontagu M (1985) Frequencies of simultaneous transformation with different T-DNAs and their relevance to the Agrobacterium plant-cell interaction. Mol Gen Genet 201: 477–484 [Google Scholar]
  • Endo M, Mikami M, Toki S (2015) Multigene knockout utilizing off-target mutations of the CRISPR/Cas9 system in rice. Plant Cell Physiol 56: 41–47 [PMC free article] [PubMed] [Google Scholar]
  • Endo M, Mikami M, Toki S (2016) Biallelic gene targeting in rice. Plant Physiol 170: 667–677 [PMC free article] [PubMed] [Google Scholar]
  • Frary A, Van Eck J (2005) Organogenesis from transformed tomato explants. Methods Mol Biol 286: 141–150 [PubMed] [Google Scholar]
  • Ghedira R, De Buck S, Van Ex F, Angenon G, Depicker A (2013) T-DNA transfer and T-DNA integration efficiencies upon Arabidopsis thaliana root explant cocultivation and floral dip transformation. Planta 238: 1025–1037 [PubMed] [Google Scholar]
  • Gómez-Gómez L, Boller T (2000) FLS2: an LRR receptor-like kinase involved in the perception of the bacterial elicitor flagellin in Arabidopsis. Mol Cell 5: 1003–1011 [PubMed] [Google Scholar]
  • Haeussler M, Schönig K, Eckert H, Eschstruth A, Mianné J, Renaud JB, Schneider-Maunoury S, Shkumatava A, Teboul L, Kent J, et al. (2016) Evaluation of off-target and on-target scoring algorithms and integration into the guide RNA selection tool CRISPOR. Genome Biol 17: 148. [PMC free article] [PubMed] [Google Scholar]
  • Hind SR, Strickler SR, Boyle PC, Dunham DM, Bao Z, O’Doherty IM, Baccile JA, Hoki JS, Viox EG, Clarke CR, et al. (2016) Tomato receptor FLAGELLIN-SENSING 3 binds flgII-28 and activates the plant immune system. Nat Plants 2: 16128. [PubMed] [Google Scholar]
  • Iaffaldano B, Zhang YX, Cornish K (2016) CRISPR/Cas9 genome editing of rubber producing dandelion Taraxacum kok-saghyz using Agrobacterium rhizogenes without selection. Ind Crops Prod 89: 356–362 [Google Scholar]
  • Jacobs TB, Martin GB (2016) High-throughput CRISPR vector construction and characterization of DNA modifications by generation of tomato hairy roots. J Vis Exp doi/10.3791/53843 [PMC free article] [PubMed] [Google Scholar]
  • Jinek M, Chylinski K, Fonfara I, Hauer M, Doudna JA, Charpentier E (2012) A programmable dual-RNA-guided DNA endonuclease in adaptive bacterial immunity. Science 337: 816–821 [PMC free article] [PubMed] [Google Scholar]
  • Kearse M, Moir R, Wilson A, Stones-Havas S, Cheung M, Sturrock S, Buxton S, Cooper A, Markowitz S, Duran C, et al. (2012) Geneious Basic: an integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics 28: 1647–1649 [PMC free article] [PubMed] [Google Scholar]
  • Krieger U, Lippman ZB, Zamir D (2010) The flowering gene SINGLE FLOWER TRUSS drives heterosis for yield in tomato. Nat Genet 42: 459–463 [PubMed] [Google Scholar]
  • Larkin PJ, Scowcroft WR (1981) Somaclonal variation: a novel source of variability from cell cultures for plant improvement. Theor Appl Genet 60: 197–214 [PubMed] [Google Scholar]
  • Lein W, Usadel B, Stitt M, Reindl A, Ehrhardt T, Sonnewald U, Börnke F (2008) Large-scale phenotyping of transgenic tobacco plants (Nicotiana tabacum) to identify essential leaf functions. Plant Biotechnol J 6: 246–263 [PubMed] [Google Scholar]
  • Liang Z, Chen K, Li T, Zhang Y, Wang Y, Zhao Q, Liu J, Zhang H, Liu C, Ran Y, et al. (2017) Efficient DNA-free genome editing of bread wheat using CRISPR/Cas9 ribonucleoprotein complexes. Nat Commun 8: 14261. [PMC free article] [PubMed] [Google Scholar]
  • Liu Y, Schiff M, Dinesh-Kumar SP (2002) Virus-induced gene silencing in tomato. Plant J 31: 777–786 [PubMed] [Google Scholar]
  • Mathieu J, Schwizer S, Martin GB (2014) Pto kinase binds two domains of AvrPtoB and its proximity to the effector E3 ligase determines if it evades degradation and activates plant immunity. PLoS Pathog 10: e1004227. [PMC free article] [PubMed] [Google Scholar]
  • McCarty DR, Settles AM, Suzuki M, Tan BC, Latshaw S, Porch T, Robin K, Baier J, Avigne W, Lai J, et al. (2005) Steady-state transposon mutagenesis in inbred maize. Plant J 44: 52–61 [PubMed] [Google Scholar]
  • Menda N, Semel Y, Peled D, Eshed Y, Zamir D (2004) In silico screening of a saturated mutation library of tomato. Plant J 38: 861–872 [PubMed] [Google Scholar]
  • Minoia S, Petrozza A, D’Onofrio O, Piron F, Mosca G, Sozio G, Cellini F, Bendahmane A, Carriero F (2010) A new mutant genetic resource for tomato crop improvement by TILLING technology. BMC Res Notes 3: 69. [PMC free article] [PubMed] [Google Scholar]
  • Miwa K, Wakuta S, Takada S, Ide K, Takano J, Naito S, Omori H, Matsunaga T, Fujiwara T (2013) Roles of BOR2, a boron exporter, in cross linking of rhamnogalacturonan II and root elongation under boron limitation in Arabidopsis. Plant Physiol 163: 1699–1709 [PMC free article] [PubMed] [Google Scholar]
  • Mladenov E, Iliakis G (2011) Induction and repair of DNA double strand breaks: the increasing spectrum of non-homologous end joining pathways. Mutat Res 711: 61–72 [PubMed] [Google Scholar]
  • Murray MG, Thompson WF (1980) Rapid isolation of high molecular weight plant DNA. Nucleic Acids Res 8: 4321–4325 [PMC free article] [PubMed] [Google Scholar]
  • Rosli HG, Zheng Y, Pombo MA, Zhong S, Bombarely A, Fei Z, Collmer A, Martin GB (2013) Transcriptomics-based screen for genes induced by flagellin and repressed by pathogen effectors identifies a cell wall-associated kinase involved in plant immunity. Genome Biol 14: R139. [PMC free article] [PubMed] [Google Scholar]
  • Saito T, Ariizumi T, Okabe Y, Asamizu E, Hiwasa-Tanase K, Fukuda N, Mizoguchi T, Yamazaki Y, Aoki K, Ezura H (2011) TOMATOMA: a novel tomato mutant database distributing Micro-Tom mutant collections. Plant Cell Physiol 52: 283–296 [PMC free article] [PubMed] [Google Scholar]
  • Schwessinger B, Ronald PC (2012) Plant innate immunity: perception of conserved microbial signatures. Annu Rev Plant Biol 63: 451–482 [PubMed] [Google Scholar]
  • Shalem O, Sanjana NE, Hartenian E, Shi X, Scott DA, Mikkelsen TS, Heckl D, Ebert BL, Root DE, Doench JG, et al. (2014) Genome-scale CRISPR-Cas9 knockout screening in human cells. Science 343: 84–87 [PMC free article] [PubMed] [Google Scholar]
  • Shalem O, Sanjana NE, Zhang F (2015) High-throughput functional genomics using CRISPR-Cas9. Nat Rev Genet 16: 299–311 [PMC free article] [PubMed] [Google Scholar]
  • Smith NA, Singh SP, Wang MB, Stoutjesdijk PA, Green AG, Waterhouse PM (2000) Total silencing by intron-spliced hairpin RNAs. Nature 407: 319–320 [PubMed] [Google Scholar]
  • Song WY, Wang GL, Chen LL, Kim HS, Pi LY, Holsten T, Gardner J, Wang B, Zhai WX, Zhu LH, et al. (1995) A receptor kinase-like protein encoded by the rice disease resistance gene, Xa21. Science 270: 1804–1806 [PubMed] [Google Scholar]
  • Svitashev S, Schwartz C, Lenderts B, Young JK, Cigan AM (2016) Genome editing in maize directed by CRISPR-Cas9 ribonucleoprotein complexes. Nat Commun 7: 13274. [PMC free article] [PubMed] [Google Scholar]
  • Tadege M, Wen J, He J, Tu H, Kwak Y, Eschstruth A, Cayrel A, Endre G, Zhao PX, Chabaud M, et al. (2008) Large-scale insertional mutagenesis using the Tnt1 retrotransposon in the model legume Medicago truncatula. Plant J 54: 335–347 [PubMed] [Google Scholar]
  • Takano J, Noguchi K, Yasumori M, Kobayashi M, Gajdos Z, Miwa K, Hayashi H, Yoneyama T, Fujiwara T (2002) Arabidopsis boron transporter for xylem loading. Nature 420: 337–340 [PubMed] [Google Scholar]
  • Torella JP, Lienert F, Boehm CR, Chen JH, Way JC, Silver PA (2014) Unique nucleotide sequence-guided assembly of repetitive DNA parts for synthetic biology applications. Nat Protoc 9: 2075–2089 [PMC free article] [PubMed] [Google Scholar]
  • Uraguchi S, Kato Y, Hanaoka H, Miwa K, Fujiwara T (2014) Generation of boron-deficiency-tolerant tomato by overexpressing an Arabidopsis thaliana borate transporter AtBOR1. Front Plant Sci 5: 125. [PMC free article] [PubMed] [Google Scholar]
  • van den Bulk RW, Löffler HJM, Lindhout WH, Koornneef M (1990) Somaclonal variation in tomato: effect of explant source and a comparison with chemical mutagenesis. Theor Appl Genet 80: 817–825 [PubMed] [Google Scholar]
  • van Kregten M, de Pater S, Romeijn R, van Schendel R, Hooykaas PJJ, Tijsterman M (2016) T-DNA integration in plants results from polymerase-θ-mediated DNA repair. Nat Plants 2: 16164. [PubMed] [Google Scholar]
  • Vazquez-Vilar M, Bernabé-Orts JM, Fernandez-Del-Carmen A, Ziarsolo P, Blanca J, Granell A, Orzaez D (2016) A modular toolbox for gRNA-Cas9 genome engineering in plants based on the GoldenBraid standard. Plant Methods 12: 10. [PMC free article] [PubMed] [Google Scholar]
  • Wang L, Albert M, Einig E, Fürst U, Krust D, Felix G (2016) The pattern-recognition receptor CORE of Solanaceae detects bacterial cold-shock protein. Nat Plants 2: 16185. [PubMed] [Google Scholar]
  • Wang T, Wei JJ, Sabatini DM, Lander ES (2014) Genetic screens in human cells using the CRISPR-Cas9 system. Science 343: 80–84 [PMC free article] [PubMed] [Google Scholar]
  • Woo JW, Kim J, Kwon SI, Corvalán C, Cho SW, Kim H, Kim SG, Kim ST, Choe S, Kim JS (2015) DNA-free genome editing in plants with preassembled CRISPR-Cas9 ribonucleoproteins. Nat Biotechnol 33: 1162–1164 [PubMed] [Google Scholar]
  • Zhang Y, Liang Z, Zong Y, Wang Y, Liu J, Chen K, Qiu JL, Gao C (2016) Efficient and transgene-free genome editing in wheat through transient expression of CRISPR/Cas9 DNA or RNA. Nat Commun 7: 12617. [PMC free article] [PubMed] [Google Scholar]
  • Zheng Y, Jiao C, Sun H, Rosli HG, Pombo MA, Zhang P, Banf M, Dai X, Martin GB, Giovannoni JJ, et al. (2016) iTAK: a program for genome-wide prediction and classification of plant transcription factors, transcriptional regulators, and protein kinases. Mol Plant 9: 1667–1670 [PubMed] [Google Scholar]
  • Zipfel C, Kunze G, Chinchilla D, Caniard A, Jones JD, Boller T, Felix G (2006) Perception of the bacterial PAMP EF-Tu by the receptor EFR restricts Agrobacterium-mediated transformation. Cell 125: 749–760 [PubMed] [Google Scholar]
  • Zipfel C, Robatzek S, Navarro L, Oakeley EJ, Jones JD, Felix G, Boller T (2004) Bacterial disease resistance in Arabidopsis through flagellin perception. Nature 428: 764–767 [PubMed] [Google Scholar]

Articles from Plant Physiology are provided here courtesy of American Society of Plant Biologists

Formats: