Hostname: page-component-848d4c4894-x5gtn Total loading time: 0 Render date: 2024-05-14T04:12:11.202Z Has data issue: false hasContentIssue false

New and emerging uses of CRISPR/Cas9 to genetically manipulate apicomplexan parasites

Published online by Cambridge University Press:  21 February 2018

Manlio Di Cristina*
Affiliation:
Department of Chemistry, Biology and Biotechnology, University of Perugia, Perugia 06122, Italy
Vern B. Carruthers
Affiliation:
Department of Microbiology and Immunology, University of Michigan Medical School, Ann Arbor, Michigan 48109, USA
*
Author for correspondence: Manlio Di Cristina, E-mail: manlio.dicristina@unipg.it or Vern B. Carruthers, E-mail: vcarruth@umich.edu

Abstract

Although the application of CRISPR/Cas9 genome engineering approaches was first reported in apicomplexan parasites only 3 years ago, this technology has rapidly become an essential component of research on apicomplexan parasites. This review briefly describes the history of CRISPR/Cas9 and the principles behind its use along with documenting its implementation in apicomplexan parasites, especially Plasmodium spp. and Toxoplasma gondii. We also discuss the recent use of CRISPR/Cas9 for whole genome screening of gene knockout mutants in T. gondii and highlight its use for seminal genetic manipulations of Cryptosporidium spp. Finally, we consider new variations of CRISPR/Cas9 that have yet to be implemented in apicomplexans. Whereas CRISPR/Cas9 has already accelerated rapid interrogation of gene function in apicomplexans, the full potential of this technology is yet to be realized as new variations and innovations are integrated into the field.

Type
Review Article
Creative Commons
Creative Common License - CCCreative Common License - BYCreative Common License - NCCreative Common License - SA
This is an Open Access article, distributed under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike licence (http://creativecommons.org/licenses/by-nc-sa/4.0/), which permits non-commercial re-use, distribution, and reproduction in any medium, provided the same Creative Commons licence is included and the original work is properly cited. The written permission of Cambridge University Press must be obtained for commercial re-use
Copyright
Copyright © Cambridge University Press 2018

Introduction

The history of the CRISPR revolution represents one of the most significant examples of basic research leading to techniques with enormous translational potential. In fact, one of the most revolutionary discoveries of the last decade took more than 25 years from the first evidence of its existence before becoming the most powerful technology for genome manipulation. CRISPR is an adaptive immune system used by several bacteria to defend themselves from infection by viruses or exogenous DNA, such as bacteriophages and plasmids, respectively. Identifying basic mechanisms underlying a viral evasion strategy may appear to be an academic interest confined to the field of microbiology. This perception, which brought frustration to many early CRISPR researchers, starting from Mojica (who first observed the presence of the CRISPR array), caused a shortage of funds, lab space and editorial rejections from leading journals (Lander, Reference Lander2016). The CRISPR odyssey parallels in some ways the discovery of another technology that radically changed molecular biology research in all organisms, the polymerase chain reaction (PCR). Description of this discovery by Kary Mullis was rejected by elite journals that missed the importance of this revolutionary finding (Campanario, Reference Campanario2009). The lesson to the non-scientific community is that basic knowledge may lead to discoveries with virtually limitless utility such as CRISPR and PCR. The biological role of CRISPR in bacterial adaptive immunity has been comprehensively described (Karginov and Hannon, Reference Karginov and Hannon2010; Wright et al. Reference Wright, Nunez and Doudna2016; Hryhorowicz et al. Reference Hryhorowicz, Lipinski, Zeyland and Slomski2017; Koonin et al. Reference Koonin, Makarova and Zhang2017; Patterson et al. Reference Patterson, Yevstigneyeva and Fineran2017). The basis of CRISPR technology, its different applications, and its potential have also been well described (Hsu et al. Reference Hsu, Lander and Zhang2014; Barrangou et al. Reference Barrangou, Birmingham, Wiemann, Beijersbergen, Hornung and Smith2015; Barrangou and Horvath, Reference Barrangou and Horvath2017; Jackson et al. Reference Jackson, McKenzie, Fagerlund, Kieper, Fineran and Brouns2017; Jiang and Doudna, Reference Jiang and Doudna2017; Kick et al. Reference Kick, Kirchner and Schneider2017; Pineda et al. Reference Pineda, Moghadam, Ebrahimkhani and Kiani2017; Salsman and Dellaire, Reference Salsman and Dellaire2017). This review instead aims to highlight the power of CRISPR technology in parasitology, principally concentrating on new advancements in apicomplexan parasites where considerable recent progress has been made.

The availability of genetic tools to manipulate genomes is crucial to understanding the biology of any organism. Within the phylum Apicomplexa, such tools have been most extensively applied to Plasmodium spp., Toxoplasma gondii and more recently Babesia spp. Research on other medically important apicomplexans such as Cryptosporidium spp., a leading cause of pathogen-induced diarrhea (Checkley et al. Reference Checkley, White, Jaganath, Arrowood, Chalmers, Chen, Fayer, Griffiths, Guerrant, Hedstrom, Huston, Kotloff, Kang, Mead, Miller, Petri, Priest, Roos, Striepen, Thompson, Ward, Van Voorhis, Xiao, Zhu and Houpt2015), has been constrained due to the lack of a long-term in vitro culture system, animal models, and molecular genetic tools, but is now surging forward with CRISPR/Cas9 technology (Vinayak et al. Reference Vinayak, Pawlowic, Sateriale, Brooks, Studstill, Bar-Peled, Cipriano and Striepen2015). Recent applications of CRISPR/Cas9 are creating exciting new opportunities to interrogate gene function and reveal important biological insight.

Adaptation of CRISPR in apicomplexan parasites has paralleled that of higher eukaryotes. The approach is based on initial generation of a double-strand DNA break (DSB) by the Cas9 nuclease in a site-specific manner driven by a single guide RNA (sgRNA) targeting an exact DNA sequence within the genome. The generation of a specific DSB activates DNA repair systems including non-homologous end joining (NHEJ), homologous repair (HR), or other alternative repair pathways, depending on the organism. NHEJ is active throughout the cell cycle, but is dominant in most organisms during G1-phase when HR is absent. NHEJ low fidelity repair of CRISPR/Cas9 induced DSBs causes deletions or insertions (indels), resulting in frame-shift mutations that typically inactivate the target gene. HR is instead an S-phase and high fidelity DNA repair pathway that can be exploited using CRISPR/Cas9 technology coupled with a ‘donor DNA’ that contains homologous sequences flanking the DSB. This HR donor DNA can be exploited to precisely insert mutations or sequences such as epitope tags, or for a complete gene knock out. Some parasites, such as Plasmodium spp. and Cryptosporidium spp. lack NHEJ and thus predominately use HR. In the absence of donor DNA Plasmodium resorts to microhomology-mediated end joining (MMEJ) where DSB is repaired using short homologous regions corresponding to as little as 4 bp flanking the lesion (Kirkman et al. Reference Kirkman, Lawrence and Deitsch2014; Singer et al. Reference Singer, Marshall, Heiss, Mair, Grimm, Mueller and Frischknecht2015). Although it usually results in a small deletion, the low frequency of MMEJ likely renders CRISPR/Cas9-mediated gene disruption in the absence of a donor DNA template highly inefficient.

This review will focus on recent applications of CRISPR/Cas9 genetic manipulation of apicomplexans, with a particular emphasis on Plasmodium spp., T. gondii and Cryptosporidium spp. Other excellent recent reviews that discuss genetic manipulation of apicomplexans in general (Suarez et al. Reference Suarez, Bishop, Alzan, Poole and Cooke2017) and CRISPR/Cas9 genome editing of protists (Lander, Reference Lander2016) also illustrate various strategies for using CRISPR/Cas9.

Plasmodium: new tools advance genetic tractability

Massive efforts have been made to develop tools to study Plasmodium spp., and particularly P. falciparum, due to the impact these parasites have on society. Malaria causes severe morbidity and mortality that is sustained by the lack of an effective malaria vaccine and the ability of the parasite to develop drug resistance. In vitro cultivation and genetic manipulation have been available to study P. falciparum for decades, but these approaches have been inefficient and time-consuming. Conventional gene knockout in P. falciparum takes months to obtain a null mutant and relies on spontaneous single- or double-crossover recombination using plasmids containing homologous sequences to the target region. Since P. falciparum lacks the machinery for NHEJ, genome integration is not random and occurs mainly in the region of homology between parasite chromosomes and plasmid DNA maintained episomally. Unfortunately, integration is a stochastic event that occurs at very low frequency and thus gene disruption in P. falciparum requires 1–3 months of continuous culture, protracted on–off cycling of drug selection, and/or negative-selection procedures. Zinc finger nuclease technology (ZFN) has been successfully used in P. falciparum to improve the generation of knockout parasites (Straimer et al. Reference Straimer, Lee, Lee, Zeitler, Williams, Pearl, Zhang, Rebar, Gregory, Llinas, Urnov and Fidock2012; Singer et al. Reference Singer, Marshall, Heiss, Mair, Grimm, Mueller and Frischknecht2015; Veiga et al. Reference Veiga, Dhingra, Henrich, Straimer, Gnadig, Uhlemann, Martin, Lehane and Fidock2016), but is limited by its targeting capability, arduous design and implementation, and high cost. ZFN vectors are also often very large and thus not suitable for genome-wide screening.

Successful establishment of CRISPR/Cas9 technology to edit Plasmodium spp. genomes has provided a powerful tool to allow rapid and efficient genetic manipulation of these parasites. Several adaptations of CRISPR/Cas9 have been developed in Plasmodium spp. based on the use of either one- or two-vector strategies, depending on whether the donor DNA is in the same or a separate plasmid as Cas9 and the sgRNA, and if different selectable markers are used to maintain the plasmids in transformed parasites. Two seminal studies in 2014 independently reported adapting CRISPR/Cas9 technology to genetically manipulate P. falciparum using different approaches to express the sgRNA (Ghorbal et al. Reference Ghorbal, Gorman, Macpherson, Martins, Scherf and Lopez-Rubio2014; Wagner et al. Reference Wagner, Platt, Goldfless, Zhang and Niles2014). In the first study, Ghorbal et al. developed a two-plasmid system expressing Cas9, under the HSP86 promoter, in one vector and the sgRNA, driven by the U6 promoter, along with the HR donor DNA flanking a selectable marker hDHFR (human dihydrofolate reductase) in a second vector. After co-transfecting P. falciparum-infected erythrocytes with both vectors, drug selection for integration of hDHFR at the target locus was applied. The sgRNA/HR donor vector also carried the negative selectable marker yfcu (yeast cytosine deaminase and uridyl phosphoribosyl transferase) to subsequently eliminate parasites carrying copies of this plasmid. This study also found that transfection of linear HR donor plasmid DNA is a viable alternative to negative selection since linear DNA can mediate recombination but does not persist (Deitsch et al. Reference Deitsch, Driskill and Wellems2001). The authors used this system to disrupt a reporter transgene (EGFP) and an endogenous gene (PfKAHRP) with hDHFR, along with introducing a point mutation in two genes (PfORC1 and PfKELCH-13) without drug selection. A possible limitation of expressing sgRNAs from the U6 promoter is that guanosine is preferred at the sgRNA 5′ position for efficient RNA polymerase III transcription. Nevertheless, the authors reported that the P. falciparum U6 promoter was able to drive expression of gRNAs in the parasite without the functional requirement for the initial guanosine nucleotide, thus expanding in P. falciparum the ability of Cas9 to target any sequence with the -NGG PAM motif. We have similarly observed in T. gondii that expression from the U6 promoter seems to be free of this initial guanosine nucleotide restriction, having successfully used sgRNAs starting with one of the other three nucleotides, although efficiency has not been determined (Di Cristina et al. Reference Di Cristina, Dou, Lunghi, Kannan, Huynh, McGovern, Schultz, Schultz, Miller, Hayes, van der Linden, Emiliani, Bogyo, Besteiro, Coppens and Carruthers2017).

In the second study, Wagner et al. also used a two-plasmid system, wherein Cas9 and the sgRNA were placed in the same vector along with a BSD (blasticidin-S deaminase) selection marker. In this system, expression of the sgRNA was driven by the T7 promoter, thus requiring expression of the T7 RNA polymerase from a second plasmid. This second plasmid also carried the HR donor and a NEO (neomycin) resistance cassette to maintain this vector episomally. The authors demonstrated disruption of two individual genes (PfKAHRP and PfEBA-175) without integration of a selectable marker.

These studies were followed by several other reports utilizing similar strategies, including some with new innovations. The Ghorbal approach allowed Nacer et al. to identify PfVAP1 (virulence-associated protein 1) as a key factor involved in the P. falciparum cytoadherence (Nacer et al. Reference Nacer, Claes, Roberts, Scheidig-Benatar, Sakamoto, Ghorbal, Lopez-Rubio and Mattei2015). Using a similar general strategy Lu et al. introduced the use of a suicide vector approach (Lu et al. Reference Lu, Tong, Pan, Yang, Liu, Tan, Zhao, Qin and Chen2016). In this work, the authors designed a suicide vector encoding Cas9 nuclease, sgRNA and a drug selection marker along with a second plasmid carrying the HR donor for cotransfection. Parasites receiving only the Cas9/sgRNA suicide vector die because of inefficient repair of the DSB by MMEJ, whereas those receiving both plasmids survive via HR repair from the donor plasmid. This approach avoids the need for a selectable marker in the donor plasmid, thus freeing up space to introduce larger knock-in tags. The system also has the potential to efficiently mediate consecutive gene manipulations. In another adaptation of the method developed by Ghorbal and colleagues, Mogollon et al. (Reference Mogollon, van Pul, Imai, Ramesar, Chevalley-Maurel, de Roo, Veld, Kroeze, Franke-Fayard, Janse and Khan2016) generated marker-free P. falciparum fluorescent reporter lines by redesigning the vector carrying the sgRNA and HR donor to include a hybrid positive and negative selectable marker (hDHFR-yfcu) outside the homology arms of the HR donor. In this new approach, positive selection is applied initially to maintain the HR donor plasmid followed by negative selection to eliminate parasites containing the episomal plasmid. In very recent work, Bryant et al. used the Ghorbal system to functionally interrogated an important conserved genetic element of var genes, the var2csa intron, without the introduction of a drug-selectable marker (Bryant et al. Reference Bryant, Regnault, Scheidig-Benatar, Baumgarten, Guizetti and Scherf2017). Also, the Wagner approach based on the sgRNA expression driven by the T7 promoter was successfully used in two other studies to introduce point mutations in either PfMDR1 (Ng et al. Reference Ng, Siciliano, Lee, de Almeida, Corey, Bopp, Bertuccini, Wittlin, Kasdin, Le Bihan, Clozel, Winzeler, Alano and Fidock2016) or PfCARL (LaMonte et al. Reference LaMonte, Lim, Wree, Reimer, Nachon, Corey, Gedeck, Plouffe, Du, Figueroa, Yeung, Bifani and Winzeler2016) to identify residues involved in drug resistance.

Distinct from the two-plasmid CRISPR/Cas9 systems developed for P. falciparum, Zhang and colleagues developed a P. yoelii system based on one-vector carrying all the products required for cleavage and DNA repair (Zhang et al. Reference Zhang, Xiao, Jiang, Zhao, Li, Gao, Ling, Wei, Li, Lu, Su, Cui and Yuan2014). The authors used this system to generate targeted deletion, reporter knock-in, and nucleotide replacement in multiple genes with high efficiency and accuracy. In subsequent work, the same group improved this single-vector system by introducing the yfcu negative selection marker to remove the episomal vector in transfected parasites after the genomic editing was achieved (Zhang et al. Reference Zhang, Gao, Yang, Jiang, Li, Wang, Xiao, Su, Cui and Yuan2017a, Reference Zhang, Gao, Wang and Zhaob). The CRISPR/Cas9 system was also employed by Knuepfer et al. to introduce a rapamycin-inducible DiCre recombinase for conditional disruption of target genes engineered with flanking loxP sites (Knuepfer et al. Reference Knuepfer, Napiorkowska, van Ooij and Holder2017). DiCre-mediated gene deletion was also achieved using the Ghorbal CRISPR/Cas9 adaptation by Volz et al. (Reference Volz, Yap, Sisquella, Thompson, Lim, Whitehead, Chen, Lampe, Tham, Wilson, Nebl, Marapana, Triglia, Wong, Rogers and Cowman2016) to demonstrate the essential role of PfRh5/PfRipr/CyRPA complex during P. falciparum invasion of erythrocytes.

Very recently, P. falciparum transfection of a purified CRISPR/Cas9-guide RNA ribonucleoprotein complex and a 200-nucleotide single-stranded oligodeoxynucleotide (ssODN) repair template introduced drug resistance mutations without the use of plasmids or the need for cloning homologous recombination templates (Crawford et al. Reference Crawford, Quan, Horst, Ebert, Wu and DeRisi2017). This approach is ideally suited for introducing mutations that confer a fitness advantage under selection conditions. The efficacy of using ssODN as a repair template in P. falciparum may greatly simplify future genome editing of these parasites.

Toxoplasma gondii: higher throughput and genome-wide screens

Contrary to Plasmodium spp., T. gondii is an apicomplexan parasite offering relative ease of growth in culture and a wide array of genetic tools that include chemical or insertional mutagenesis, homologous gene replacement, conditional knockdown and tagging techniques (Wang et al. Reference Wang, Huang, Behnke, Chen, Shen and Zhu2016a). The availability of numerous selectable markers for generation of stable Toxoplasma strains (Roos et al. Reference Roos, Donald, Morrissette and Moulton1994, Reference Roos, Sullivan, Striepen, Bohne and Donald1997; Donald and Roos, Reference Donald and Roos1995, Reference Donald and Roos1998; Fox et al. Reference Fox, Belperron and Bzik1999, Reference Fox, Belperron and Bzik2001; Soete et al. Reference Soete, Hettman and Soldati1999; Wang et al. Reference Wang, Huang, Behnke, Chen, Shen and Zhu2016a) and the rapid loss of exogenous non-integrated DNA, with no detectable exogenous DNA 7 days post-transfection (Soldati and Boothroyd, Reference Soldati and Boothroyd1993; Black and Boothroyd, Reference Black and Boothroyd1998), makes the study of the biology of T. gondii by genetic manipulations more manageable. Integration of foreign DNA in T. gondii is relatively efficient, with recombination rates of ~0.1% without restriction enzyme-mediated integration (REMI) and ~5% with REMI (Black et al. Reference Black, Seeber, Soldati, Kim and Boothroyd1995; Roos et al. Reference Roos, Sullivan, Striepen, Bohne and Donald1997). The presence of a very active NHEJ pathway in T. gondii is a limitation for generating strains with homology-directed knockout or the precise insertion of tags or mutations because of a high prevalence of random DNA integration. Due to this feature, conventional generation of knock out strains for T. gondii is inefficient and requires homology flanks of 2–3 kbp (Donald and Roos, Reference Donald and Roos1994; Roos et al. Reference Roos, Sullivan, Striepen, Bohne and Donald1997; Zhang et al. Reference Zhang, Kim, Ma, Wittner, Tanowitz and Weiss1999; Craver and Knoll, Reference Craver and Knoll2007). To overcome this, T. gondii type I (e.g., RH) or type II (e.g., Prugniaud, Pru) strains were modified to disrupt NHEJ-mediated insertion by deleting one key component of this pathway, the gene encoding the Ku80 protein (Fox et al. Reference Fox, Ristuccia, Gigley and Bzik2009, Reference Fox, Falla, Rommereim, Tomita, Gigley, Mercier, Cesbron-Delauw, Weiss and Bzik2011; Huynh and Carruthers, Reference Huynh and Carruthers2009). The impact of these two strains, named RHΔku80 and PruΔku80, on understanding Toxoplasma biology has been substantial, allowing higher fidelity, rapid generation of knockout strains, and the introduction of epitope tags. Although the generation of Δku80 strains minimized the problem of random DNA integration, this approach, of course, limits the studies to these two strains currently. Moreover, since the Ku80 protein is involved in DNA repair, strains lacking this gene may be prone to accumulate genetic mutations after prolonged culture.

Introduction of CRISPR/Cas9 into the Toxoplasma field has revolutionized the capability to efficiently generate gene knockouts in any strain. Several CRISPR/Cas9-based approaches have been developed in different laboratories to inactivate selected gene function. For example, complete or partial deletion of the target sequence was obtained through double crossover triggered by a site-directed sgRNA/Cas9-mediated DSB and subsequent DNA repair using a donor DNA comprised a drug resistance expression cassette flanked by about 1 kbp of DNA homologous to the target locus (Shen et al. Reference Shen, Brown, Lee and Sibley2014a). This approach also substantially increased the throughput of gene deletions, exemplified by individual or sequential disruption of entire gene families (Shen et al. Reference Shen, Buguliskis, Lee and Sibley2014b). Gene inactivation was also obtained via ‘indels’ in the coding region generated by NHEJ repair of a sgRNA/Cas9-mediated DSB (Sidik et al. Reference Sidik, Hackett, Tran, Westwood and Lourido2014; Wang et al. Reference Wang, Huang, Li, Chen, Ning and Zhu2016b). A ‘clean’ knockout with complete gene deletion is desirable because it avoids the potential expression of a truncated protein and precludes homologous reinsertion of the gene or cDNA for genetic complementation. Nonetheless, template mediated complete gene deletion is less efficient in non-Δku80 strains because homologous recombination is active only when parasites are in the S/G2 phase and thus NHEJ is the prevalent form of DNA repair in T. gondii when extracellular, G0-phase parasites are used for transfection. To enhance HR vs. NHEJ events, Behnke and colleagues (Behnke et al. Reference Behnke, Khan, Lauron, Jimah, Wang, Tolia and Sibley2015) developed a new CRISPR vector that expresses Cas9 and two sgRNAs. The two sgRNAs direct Cas9 to generate two DSBs, one at each end of the target gene or locus. This tactic not only improves efficiency, but it also permits efficient disruption of large genes or tandem gene arrays. In our hands, the two-sgRNA approach allowed successful knock out of several genes individually including those encoding TgCPL (cathepsin protease L) (Di Cristina et al. Reference Di Cristina, Dou, Lunghi, Kannan, Huynh, McGovern, Schultz, Schultz, Miller, Hayes, van der Linden, Emiliani, Bogyo, Besteiro, Coppens and Carruthers2017) and TgASP1 (aspartic protease 1) (Di Cristina and Carruthers, unpublished) in the T. gondii type II strain ME49. In summary, CRISPR/Cas9 is opening up genetic manipulation of any T. gondii strain and allowing for large genetic disruptions to interrogate gene function.

Although CRISPR/Cas9 is versatile for any strain, applying it in a Δku80 strain provides the advantage of using short homology sequences, thereby permitting convenient and precise gene knockouts or knocking of tags or mutations. For example, Sidik and co-authors (Sidik et al. Reference Sidik, Hackett, Tran, Westwood and Lourido2014) introduced tags or mutations in a Δku80 strain using synthetic oligonucleotide repair templates with 40 bp of homology without the need for a selectable marker. This strategy is based on co-transfecting a Cas9 + sgRNA expression plasmid with a synthetic double strand oligonucleotide bearing the desired tag or mutation together with a silent mutation to eliminate the PAM site NGG beside the 20 bp sequence targeted by the gRNA. The efficiency of this approach was enhanced by fluorescence-activated cell sorting (FACS) parasites that received the Cas9/sgRNA vector, exploiting the fluorescence emitted by the GFP fused to the Cas9 protein. Only ~20–30% of parasites obtain the Cas9/sgRNA vector after electroporation, making it critical to remove the predominant fraction of non-transfected parasites to enrich the population with edited parasites. Since FACS is expensive and not available for all the laboratories, we developed a protocol that allows the enrichment of Cas9-expressing parasites with a Cas9/sgRNA vector bearing a bleomycin resistance gene (Di Cristina and Carruthers, unpublished). Parasites receiving this new plasmid, named pCas9/sgRNA/Bleo, are subjected to phleomycin treatment 24 h after transfection to eliminate parasites that have not received the plasmid and enrich for parasites expressing Cas9 and the sgRNA. Treatment one day after transfection ensures that bleomycin resistance is transiently expressed by the 20–30% of the population that incorporate the plasmid. Stable integration is not favoured due to the Cas9 toxicity, which works as a negative selection against vector integration. In our hands, this approach results in about 30–80% efficiency of edited or tagged parasites, depending on the impact to parasite fitness of the mutation introduced, allowing easy identification of single mutant clones. Beyond this variation of the approach, the reader is referred to an excellent how-to guide for using CRISPR/Cas9 for various applications in Toxoplasma that was published recently (Shen et al. Reference Shen, Brown, Long and Sibley2017).

High-throughput strategies to genetically engineer and screen large numbers of mutants or populations are powerful weapons in modern systems biology. To this end, the emergence of CRISPR/Cas9 has prompted new screening approaches using sgRNA libraries to perform genome-wide knockouts in a parasite population. Such approaches allow measuring the fitness contribution of every gene in the parasite genome. Exploiting the high rates of NHEJ in T. gondii, Sebastian Lourido's laboratory efficiently created frame-shift mutations and insertions at the DSBs generated by transfecting a Cas9-expressing RH strain with a library of sgRNAs containing 10 guides against each of the 8158 predicted T. gondii protein-coding genes (Sidik et al. Reference Sidik, Huet, Ganesan, Huynh, Wang, Nasamu, Thiru, Saeij, Carruthers, Niles and Lourido2016). The guide RNA library was cloned into the sgRNA expression vector and the integrated sgRNAs were exploited as barcodes to measure the contribution of each gene to parasite fitness. Generation of a Cas9-expressing RH strain was likely instrumental in obtaining high rates of gene disruption. Due to the toxicity of Cas9 expression in T. gondii, as observed for other microorganisms (Jiang et al. Reference Jiang, Brueggeman, Horken, Plucinak and Weeks2014; Peng et al. Reference Peng, Kurup, Yao, Minning and Tarleton2014), the authors developed a strategy to obtain strains of T. gondii stably expressing this nuclease by co-expressing a decoy sgRNA to prevent the detrimental effect to parasites by unintended Cas9 activity directed by endogenous RNAs. This work represents the first genome-wide functional analysis of an apicomplexan, thus providing broad-based functional information on T. gondii genes and their contributions to parasite fitness during infection of human fibroblasts. One initial limitation of this outstanding work is its restriction to the tachyzoite stage of T. gondii, the rapidly growing form responsible for the acute phase of the infection. Applying this approach to other life stages will require improvements to the transfection and integration efficiencies of strains that competently differentiate into other stages. The contribution to tachyzoite fitness of each predicted T. gondii protein-coding genes is now available (www.toxodb.org).

Recently, David Sibley's group developed an auxin-inducible degron (AID) tagging system for conditional protein depletion in T. gondii (Brown et al. Reference Brown, Long and Sibley2017; Long et al. Reference Long, Brown, Drewry, Anthony, Phan and Sibley2017). They exploited a new combination of CRISPR/Cas9-mediated gene editing and a plant-derived AID system to identify which cyclic GMP (cGMP)-dependent protein kinase G (PKG) isoforms are necessary for PKG-dependent cellular processes (Brown et al. Reference Brown, Long and Sibley2017) and to examine the roles of three apically localized calmodululin-like proteins (Long et al. Reference Long, Brown, Drewry, Anthony, Phan and Sibley2017). Adaptation of the AID system to T. gondii adds a powerful new tool to identify the consequences of rapidly down-regulating expression of cytosolic proteins to infer function.

CRISPR technology has been adapted to apicomplexan parasites relatively recently and thus has not been fully exploited and expanded. In mammals, evolution of this technology led to the development of tissue or time-specific promoters to restrict the genome editing to a precise cell type or developmental stage (Harrison et al. Reference Harrison, Jenkins, O'Connor-Giles and Wildonger2014; Ablain et al. Reference Ablain, Durand, Yang, Zhou and Zon2015; Bortesi and Fischer, Reference Bortesi and Fischer2015; Wang et al. Reference Wang, Li, Zhao, Li, Zhou and Liu2015; Yoshioka et al. Reference Yoshioka, Fujii, Ogawa, Sugiura and Naito2015; Lee et al. Reference Lee, Ng and Ingham2016; Xu et al. Reference Xu, Zhao, Gao, Xu and Han2017; Zhang et al. Reference Zhang, Gao, Wang and Zhao2017b). In the classic CRISPR technology, sgRNAs are usually transcribed under control of RNA polymerase III promoters to obtain transcripts devoid of both capping and poly(A) tails, thereby generating the correct 5′-end of the sgRNA and avoiding exportation of the sgRNA to the cytoplasm, respectively. Tissue/developmental-specific sgRNA expression requires using RNA polymerase II promoters active exclusively in the desired cell type or stage. To generate RNA polymerase II-driven functional sgRNAs, a strategy based on the use of ribozymes was developed by several laboratories (Yoshioka et al. Reference Yoshioka, Fujii, Ogawa, Sugiura and Naito2015; Ng and Dean, Reference Ng and Dean2017; Xu et al. Reference Xu, Zhao, Gao, Xu and Han2017; Zhang et al. Reference Zhang, Gao, Yang, Jiang, Li, Wang, Xiao, Su, Cui and Yuan2017a, Reference Zhang, Gao, Wang and Zhaob). Hammerhead or hepatitis delta virus ribozymes perform site-specific self-cleavage, resulting in mature sgRNAs with correct 5′- and 3′-ends. This ribozyme-flanked gRNA expression system can be exploited for the spatiotemporal expression of gRNA employing cell type or developmental-specific promoters. In Toxoplasma, this approach has not been investigated yet but may allow programmed gene inactivation in a stage-specific manner by exploiting promoters active at specific phases of the parasite life cycle. For example expressing the ribozyme-sgRNA and Cas9 under the control of a bradyzoite-specific promoter could allow stage-specific inactivation of all genes, even those that are essential for tachyzoites. This approach would allow assessing the role of essential genes during the chronic stage of T. gondii. Moreover, multiple sgRNAs linked with self-cleaving ribozymes could be simultaneously expressed from a single promoter to exert genome editing at different sites. Alternatively, several new approaches permit conditional expression of Cas9 via chemical or optical activation (Nihongaki et al. Reference Nihongaki, Kawano, Nakajima and Sato2015; Polstein and Gersbach, Reference Polstein and Gersbach2015; Wright et al. Reference Wright, Sternberg, Taylor, Staahl, Bardales, Kornfeld and Doudna2015; Zetsche et al. Reference Zetsche, Volz and Zhang2015; Liu et al. Reference Liu, Ramli, Woo, Wang, Zhao, Zhang, Yim, Chong, Gowher, Chua, Jung, Lee and Tan2016). In principle, it should also be possible to append a destabilization domain to Cas9 for ligand-dependent expression of Cas9 in any stage of the parasite.

Cryptosporidium: introducing a new genetic era

Cryptosporidium spp. causes severe diarrhoea in young children, with 10% mortality in such cases (Liu et al. Reference Liu, Johnson, Cousens, Perin, Scott, Lawn, Rudan, Campbell, Cibulskis and Li2012). Cryptosporidiosis also causes life-threatening chronic disease in immunocompromised individuals, including those afflicted by HIV/AIDS. Infections occur worldwide in association with oocyst contaminated water, with no vaccines available, and only a single drug (nitazoxanide) has been approved with limited benefit for malnourished children and immunocompromised patients (Amadi et al. Reference Amadi, Mwiya, Musuku, Watuka, Sianongo, Ayoub and Kelly2002, Reference Amadi, Mwiya, Sianongo, Payne, Watuka, Katubulushi and Kelly2009). Progress in understanding Cryptosporidium spp. biology and developing new treatments have been hindered by the limited tractability of the parasite, which includes a lack of systems for continuous culture, the absence of facile animal models, and the dearth of molecular genetic tools (Striepen, Reference Striepen2013; Checkley et al. Reference Checkley, White, Jaganath, Arrowood, Chalmers, Chen, Fayer, Griffiths, Guerrant, Hedstrom, Huston, Kotloff, Kang, Mead, Miller, Petri, Priest, Roos, Striepen, Thompson, Ward, Van Voorhis, Xiao, Zhu and Houpt2015). Cryptosporidium spp. cultures last a few days in vitro since parasites undergo one or two rounds of replication at most, limiting experiments to small numbers of parasites during a fraction of the life cycle. Species that infect humans cannot be easily studied in standard model hosts such as mice. Also, since Cryptosporidium spp. is intrinsically refractory to antifolate drugs, selection of genetically modified parasites using these drugs, as for Toxoplasma and Plasmodium spp., is not possible. Further hindering Cryptosporidium spp. genomic manipulation, transient transfection is 10 000-fold less efficient than that of T. gondii (Vinayak et al. Reference Vinayak, Pawlowic, Sateriale, Brooks, Studstill, Bar-Peled, Cipriano and Striepen2015). Genetic validation of potential drug targets for Cryptosporidium spp. is a key unmet need.

Toward this goal, CRISPR/Cas9 technology has proven again to be a powerful system even for Cryptosporidium spp. Boris Striepen's laboratory recently developed for the first time a protocol for transfecting C. parvum sporozoites in tissue culture and isolation of stable genetically modified parasites (Vinayak et al. Reference Vinayak, Pawlowic, Sateriale, Brooks, Studstill, Bar-Peled, Cipriano and Striepen2015). Notwithstanding a 10-fold optimization of transfection, the low efficiency still required the use of a highly sensitive nano-luciferase (nLuc) reporter. Sporozoite expression of nLuc was achieved using the strong enolase promoter, whilst the aminoglycoside antibiotic paromomycin was used as selection marker since it is effective in tissue culture and in immunocompromised mice (Theodos et al. Reference Theodos, Griffiths, D'Onfro, Fairfield and Tzipori1998). After electroporation, transfected sporozoites were directly introduced into the mouse intestine by surgery due to the low oral infection efficiency of this stage. Expression of sgRNA and the Cas9 nuclease was achieved using the C. parvum U6 and aldolase promoters, respectively. Since Cryptosporidium spp. lack NHEJ, similar to Plasmodium species, transgene integration is likely to require homologous recombination. Thus, in a series of elegant experiments, the Striepen group restored a dead version of the nLuc carrying a stop codon that ablated luciferase activity by using short double-stranded templates for repair. This demonstrated that genome editing through homologous recombination was also possible in the recalcitrant Cryptosporidium spp. The authors also achieved for the first time a gene knock out in this parasite by deleting the gene encoding thymidine kinase (TK). This genetic manipulation provided evidence of the non-essentiality of TK and its role as an alternative route for thymidine monophosphate synthesis, explaining why C. parvum tolerates high doses of antifolate drugs. The Striepen group also recently published a how-to guide for genetic manipulation of Cryptosporidium spp. that will be invaluable to the field (Pawlowic et al. Reference Pawlowic, Vinayak, Sateriale, Brooks and Striepen2017). In summary, an adaptation of CRISPR/Cas9 technology to this nearly intractable parasite allowed generation of the first Cryptosporidium knock out strain and, at the same time, deletion of the TK gene provided a new potential selection marker for genome manipulation. Overcoming such barriers for Cryptosporidium spp. opens new avenues to import the RNA- or protein-based regulatory strategies developed from other apicomplexans.

Future directions

Another key aspect of CRISPR is the impact of off-target effects that may introduce breaks in genomic sites other than the specific sgRNA target. Cas9 tolerates mismatches between guide RNA and target DNA differently depending on the position of the mismatches. Mismatches are tolerated at the 5′-end of the target site, but not at the 3′-end ‘seed’ sequence beside the PAM (Semenova et al. Reference Semenova, Jore, Datsenko, Semenova, Westra, Wanner, van der Oost, Brouns and Severinov2011; Cho et al. Reference Cho, Kim, Kim and Kim2013; Cong et al. Reference Cong, Ran, Cox, Lin, Barretto, Habib, Hsu, Wu, Jiang, Marraffini and Zhang2013; Ma et al. Reference Ma, Zhang and Huang2014; Farboud and Meyer, Reference Farboud and Meyer2015; Port and Bullock, Reference Port and Bullock2016). Although the introduction of unwanted changes in sequences of the genome may cause unpredictable consequences for the parasite phenotype, this might not be a major concern for apicomplexans because their small genomes make off-target mutations less likely. No evidence of off-target mutations introduced by Cas9 was seen in both P. falciparum (Ghorbal et al. Reference Ghorbal, Gorman, Macpherson, Martins, Scherf and Lopez-Rubio2014; Wagner et al. Reference Wagner, Platt, Goldfless, Zhang and Niles2014) and P. yoelii (Zhang et al. Reference Zhang, Xiao, Jiang, Zhao, Li, Gao, Ling, Wei, Li, Lu, Su, Cui and Yuan2014), suggesting that this system is very specific in these parasites lacking NHEJ. Off-target effects have not been fully explored in other apicomplexan parasites, such as T. gondii, that have an active NHEJ system and thus may also repair DSBs in off-target positions. Regardless, the recent development of high-fidelity Cas9 variants may be useful in apicomplexans to minimize off-target effects (Kleinstiver et al. Reference Kleinstiver, Pattanayak, Prew, Tsai, Nguyen, Zheng and Joung2016; Slaymaker et al. Reference Slaymaker, Gao, Zetsche, Scott, Yan and Zhang2016).

Recently, the potential of the CRISPR/Cas9 system has been further expanded to regulate transcription or introduce epigenetic modification in target genes. Activation (CRISPRa) or Repression (CRISPRi) of transcription of target genes has been achieved with a catalytically inactive Cas9 protein (dCas9) lacking endonuclease activity fused to activating or repressive effectors. These systems have the advantage of controlling gene expression in an inducible and reversible manner (Qi et al. Reference Qi, Larson, Gilbert, Doudna, Weissman, Arkin and Lim2013). CRISPR/Cas9-directed epigenetic modifications were achieved by fusing the dCas9 protein to epigenetic effectors (e.g., DNA demethylase, histone acetyltransferase and others) for epigenomic engineering (Hilton et al. Reference Hilton, D'Ippolito, Vockley, Thakore, Crawford, Reddy and Gersbach2015; Kearns et al. Reference Kearns, Pham, Tabak, Genga, Silverstein, Garber and Maehr2015). The CRISPR/Cas9 system has also been adapted to cleave single-stranded RNA at specific target sites by providing a PAM as part of an oligonucleotide (PAMmer) that hybridizes to the target RNA. In this way, an RNA-targeting Cas9 protein (RCas9) was directed to bind and cleave target RNAs at specific sites using specially designed PAMmers, enabling specific RNA degradation (O'Connell et al. Reference O'Connell, Oakes, Sternberg, East-Seletsky, Kaplan and Doudna2014). Since apicomplexans lack or have an incomplete system for RNA interference, this strategy might represent an alternative to RNA silencing. A further application of the CRISPR/Cas9 technology has been recently developed to study topologically associated domains (TADs), i.e. genome organization of chromatin into ordered and hierarchical topological structures in interphase nuclei (Bouwman and de Laat, Reference Bouwman and de Laat2015; Sexton and Cavalli, Reference Sexton and Cavalli2015; Bonev and Cavalli, Reference Bonev and Cavalli2016). TADs play important roles in various nuclear processes such as gene regulation since distal elements regulate their gene targets through specific chromatin-looping contacts such as long-distance enhancer–promoter interactions. CRISPR/Cas9 technology provides great opportunities to study TADs by probing spatial DNA–looping interactions and perturb higher-order chromatin organization (Huang and Wu, Reference Huang and Wu2016). TADs have been poorly characterized in apicomplexan parasites and thus this new CRISPR/Cas9 approach offers fresh tools to better understand chromatin organization, opening new avenues to understanding the evolution of chromatin organization from unicellular to multicellular organisms.

Acknowledgements

We thank My-Hang (Mae) Huynh and Aric J. Schultz for critically reading the manuscript.

Financial support

This work was supported by the U.S. National Institutes of Health (V.B.C., grant number R01AI046675) (V.B.C. and M.D.C, grant number R01AI120607).

References

Ablain, J, Durand, EM, Yang, S, Zhou, Y and Zon, LI (2015) A CRISPR/Cas9 vector system for tissue-specific gene disruption in zebrafish. Developmental Cell 32, 756764.Google Scholar
Amadi, B, Mwiya, M, Musuku, J, Watuka, A, Sianongo, S, Ayoub, A and Kelly, P (2002) Effect of nitazoxanide on morbidity and mortality in Zambian children with cryptosporidiosis: a randomised controlled trial. The Lancet 360, 13751380.Google Scholar
Amadi, B, Mwiya, M, Sianongo, S, Payne, L, Watuka, A, Katubulushi, M and Kelly, P (2009) High dose prolonged treatment with nitazoxanide is not effective for cryptosporidiosis in HIV positive Zambian children: a randomised controlled trial. BMC Infectious Diseases 9, 195.Google Scholar
Barrangou, R and Horvath, P (2017) A decade of discovery: CRISPR functions and applications. Nature Microbiology 2, 17092.Google Scholar
Barrangou, R, Birmingham, A, Wiemann, S, Beijersbergen, RL, Hornung, V and Smith, A (2015) Advances in CRISPR-Cas9 genome engineering: lessons learned from RNA interference. Nucleic Acids Research 43, 34073419.Google Scholar
Behnke, MS, Khan, A, Lauron, EJ, Jimah, JR, Wang, Q, Tolia, NH and Sibley, LD (2015) Rhoptry proteins ROP5 and ROP18 are major murine virulence factors in genetically divergent South American strains of Toxoplasma gondii. PLoS Genetics 11, e1005434.Google Scholar
Black, M, Seeber, F, Soldati, D, Kim, K and Boothroyd, JC (1995) Restriction enzyme-mediated integration elevates transformation frequency and enables co-transfection of Toxoplasma gondii. Molecular and Biochemical Parasitology 74, 5563.Google Scholar
Black, MW and Boothroyd, JC (1998) Development of a stable episomal shuttle vector for Toxoplasma gondii. Journal of Biological Chemistry 273, 39723979.Google Scholar
Bonev, B and Cavalli, G (2016) Organization and function of the 3D genome. Nature Reviews Genetics 17, 661678.Google Scholar
Bortesi, L and Fischer, R (2015) The CRISPR/Cas9 system for plant genome editing and beyond. Biotechnology Advances 33, 4152.Google Scholar
Bouwman, BA and de Laat, W (2015) Architectural hallmarks of the pluripotent genome. FEBS Letters 589, 29052913.Google Scholar
Brown, KM, Long, S and Sibley, LD (2017) Plasma membrane association by N-acylation governs PKG function in Toxoplasma gondii. mBio 8, e0037517.Google Scholar
Bryant, JM, Regnault, C, Scheidig-Benatar, C, Baumgarten, S, Guizetti, J and Scherf, A (2017) CRISPR/cas9 genome editing reveals that the intron is not essential for var2csa gene activation or silencing in Plasmodium falciparum. mBio 8, e0072917.Google Scholar
Campanario, J (2009) Rejecting and resisting Nobel class discoveries: accounts by Nobel Laureates. Scientometrics 81, 549565.Google Scholar
Checkley, W, White, AC Jr., Jaganath, D, Arrowood, MJ, Chalmers, RM, Chen, XM, Fayer, R, Griffiths, JK, Guerrant, RL, Hedstrom, L, Huston, CD, Kotloff, KL, Kang, G, Mead, JR, Miller, M, Petri, WA Jr., Priest, JW, Roos, DS, Striepen, B, Thompson, RC, Ward, HD, Van Voorhis, WA, Xiao, L, Zhu, G and Houpt, ER (2015) A review of the global burden, novel diagnostics, therapeutics, and vaccine targets for cryptosporidium. The Lancet Infectious Diseases 15, 8594.Google Scholar
Cho, SW, Kim, S, Kim, JM and Kim, J (2013) Targeted genome engineering in human cells with the Cas9 RNA-guided endonuclease. Nature Biotechnology 31, 230232.Google Scholar
Cong, L, Ran, FA, Cox, D, Lin, S, Barretto, R, Habib, N, Hsu, PD, Wu, X, Jiang, W, Marraffini, LA and Zhang, F (2013) Multiplex genome engineering using CRISPR/Cas systems. Science (New York, NY) 339, 819823.Google Scholar
Craver, MP and Knoll, LJ (2007) Increased efficiency of homologous recombination in Toxoplasma gondii dense granule protein 3 demonstrates that GRA3 is not necessary in cell culture but does contribute to virulence. Molecular and Biochemical Parasitology 153, 149157.Google Scholar
Crawford, ED, Quan, J, Horst, JA, Ebert, D, Wu, W and DeRisi, JL (2017) Plasmid-free CRISPR/Cas9 genome editing in Plasmodium falciparum confirms mutations conferring resistance to the dihydroisoquinolone clinical candidate SJ733. PLoS ONE 12, e0178163.Google Scholar
Deitsch, K, Driskill, C and Wellems, T (2001) Transformation of malaria parasites by the spontaneous uptake and expression of DNA from human erythrocytes. Nucleic Acids Research 29, 850853.Google Scholar
Di Cristina, M, Dou, Z, Lunghi, M, Kannan, G, Huynh, MH, McGovern, OL, Schultz, TL, Schultz, AJ, Miller, AJ, Hayes, BM, van der Linden, W, Emiliani, C, Bogyo, M, Besteiro, S, Coppens, I and Carruthers, VB (2017) Toxoplasma depends on lysosomal consumption of autophagosomes for persistent infection. Nature Microbiology 2, 17096.Google Scholar
Donald, RG and Roos, DS (1998) Gene knock-outs and allelic replacements in Toxoplasma gondii: HXGPRT as a selectable marker for hit-and-run mutagenesis. Molecular & Biochemical Parasitology 91, 295305.Google Scholar
Donald, RGK and Roos, DS (1994) Homologous recombination and gene replacement at the dihydrofolate reductase-thymidylate synthase locus in Toxoplasma gondii. Molecular & Biochemical Parasitology 63, 243253.Google Scholar
Donald, RGK and Roos, DS (1995) Insertional mutagenesis and marker rescue in a protozoan parasite: cloning of the uracil phosphoribosyltransferase locus from Toxoplasma gondii. Proceedings of the National Academy of Sciences USA 92, 57495753.Google Scholar
Farboud, B and Meyer, BJ (2015) Dramatic enhancement of genome editing by CRISPR/Cas9 through improved guide RNA design. Genetics 199, 959971.Google Scholar
Fox, BA, Belperron, AA and Bzik, DJ (1999) Stable transformation of Toxoplasma gondii based on a pyrimethamine resistant trifunctional dihydrofolate reductase-cytosine deaminase-thymidylate synthase gene that confers sensitivity to 5-fluorocytosine. Molecular & Biochemical Parasitology 98, 93103.Google Scholar
Fox, BA, Belperron, AA and Bzik, DJ (2001) Negative selection of herpes simplex virus thymidine kinase in Toxoplasma gondii. Molecular & Biochemical Parasitology 116, 8588.Google Scholar
Fox, BA, Ristuccia, JG, Gigley, JP and Bzik, DJ (2009) Efficient gene replacements in Toxoplasma gondii strains deficient for nonhomologous end-joining. Eukaryotic Cell 8, 520529.Google Scholar
Fox, BA, Falla, A, Rommereim, LM, Tomita, T, Gigley, JP, Mercier, C, Cesbron-Delauw, MF, Weiss, LM and Bzik, DJ (2011) Type II Toxoplasma gondii KU80 knockout strains enable functional analysis of genes required for cyst development and latent infection. Eukaryotic Cell 10, 11931206.Google Scholar
Ghorbal, M, Gorman, M, Macpherson, CR, Martins, RM, Scherf, A and Lopez-Rubio, JJ (2014) Genome editing in the human malaria parasite Plasmodium falciparum using the CRISPR-Cas9 system. Nature Biotechnology 32, 819821.Google Scholar
Harrison, MM, Jenkins, BV, O'Connor-Giles, KM and Wildonger, J (2014) A CRISPR view of development. Genes & Development 28, 18591872.Google Scholar
Hilton, IB, D'Ippolito, AM, Vockley, CM, Thakore, PI, Crawford, GE, Reddy, TE and Gersbach, CA (2015) Epigenome editing by a CRISPR-Cas9-based acetyltransferase activates genes from promoters and enhancers. Nature Biotechnology 33, 510517.Google Scholar
Hryhorowicz, M, Lipinski, D, Zeyland, J and Slomski, R (2017) CRISPR/cas9 immune system as a tool for genome engineering. Archivum Immunologiae et Therapiae Experimentalis 65, 233240.Google Scholar
Hsu, PD, Lander, ES and Zhang, F (2014) Development and applications of CRISPR-Cas9 for genome engineering. Cell 157, 12621278.Google Scholar
Huang, H and Wu, Q (2016) CRISPR double cutting through the labyrinthine architecture of 3D genomes. Journal of Genetics and Genomics = Yi chuan xue bao 43, 273288.Google Scholar
Huynh, MH and Carruthers, VB (2009) Tagging of endogenous genes in a Toxoplasma gondii strain lacking Ku80. Eukaryotic Cell 8, 530539.Google Scholar
Jackson, SA, McKenzie, RE, Fagerlund, RD, Kieper, SN, Fineran, PC and Brouns, SJ (2017) CRISPR-Cas: adapting to change. Science (New York, NY) 356, eaal5056. Epub 2017 Apr 6.Google Scholar
Jiang, F and Doudna, JA (2017) CRISPR-Cas9 structures and mechanisms. Annual Review of Biophysics 46, 505529.Google Scholar
Jiang, W, Brueggeman, AJ, Horken, KM, Plucinak, TM and Weeks, DP (2014) Successful transient expression of Cas9 and single guide RNA genes in Chlamydomonas reinhardtii. Eukaryotic Cell 13, 14651469.Google Scholar
Karginov, FV and Hannon, GJ (2010) The CRISPR system: small RNA-guided defense in bacteria and archaea. Molecular Cell 37, 719.Google Scholar
Kearns, NA, Pham, H, Tabak, B, Genga, RM, Silverstein, NJ, Garber, M and Maehr, R (2015) Functional annotation of native enhancers with a Cas9-histone demethylase fusion. Nature Methods 12, 401403.Google Scholar
Kick, L, Kirchner, M and Schneider, S (2017) CRISPR-Cas9: from a bacterial immune system to genome-edited human cells in clinical trials. Bioengineered 8, 280286.Google Scholar
Kirkman, LA, Lawrence, EA and Deitsch, KW (2014) Malaria parasites utilize both homologous recombination and alternative end joining pathways to maintain genome integrity. Nucleic Acids Research 42, 370379.Google Scholar
Kleinstiver, BP, Pattanayak, V, Prew, MS, Tsai, SQ, Nguyen, NT, Zheng, Z and Joung, JK (2016) High-fidelity CRISPR-Cas9 nucleases with no detectable genome-wide off-target effects. Nature 529, 490495.Google Scholar
Knuepfer, E, Napiorkowska, M, van Ooij, C and Holder, AA (2017) Generating conditional gene knockouts in Plasmodium – a toolkit to produce stable DiCre recombinase-expressing parasite lines using CRISPR/Cas9. Scientific Reports 7, 3881-017-03984-3.Google Scholar
Koonin, EV, Makarova, KS and Zhang, F (2017) Diversity, classification and evolution of CRISPR-Cas systems. Current Opinion in Microbiology 37, 6778.Google Scholar
LaMonte, G, Lim, MY, Wree, M, Reimer, C, Nachon, M, Corey, V, Gedeck, P, Plouffe, D, Du, A, Figueroa, N, Yeung, B, Bifani, P and Winzeler, EA (2016) Mutations in the Plasmodium falciparum cyclic amine resistance locus (PfCARL) confer multidrug resistance. mBio 7, e00696-16.Google Scholar
Lander, ES (2016) The heroes of CRISPR. Cell 164, 1828.Google Scholar
Lee, RT, Ng, AS and Ingham, PW (2016) Ribozyme mediated gRNA generation for in vitro and in vivo CRISPR/Cas9 mutagenesis. PLoS ONE 11, e0166020.Google Scholar
Liu, KI, Ramli, MN, Woo, CW, Wang, Y, Zhao, T, Zhang, X, Yim, GR, Chong, BY, Gowher, A, Chua, MZ, Jung, J, Lee, JH and Tan, MH (2016) A chemical-inducible CRISPR-Cas9 system for rapid control of genome editing. Nature Chemical Biology 12, 980987.Google Scholar
Liu, L, Johnson, HL, Cousens, S, Perin, J, Scott, S, Lawn, JE, Rudan, I, Campbell, H, Cibulskis, R and Li, M (2012) Global, regional, and national causes of child mortality: an updated systematic analysis for 2010 with time trends since 2000. The Lancet 379, 21512161.Google Scholar
Long, S, Brown, KM, Drewry, LL, Anthony, B, Phan, IQH and Sibley, LD (2017) Calmodulin-like proteins localized to the conoid regulate motility and cell invasion by Toxoplasma gondii. PLoS Pathogens 13, e1006379.Google Scholar
Lu, J, Tong, Y, Pan, J, Yang, Y, Liu, Q, Tan, X, Zhao, S, Qin, L and Chen, X (2016) A redesigned CRISPR/Cas9 system for marker-free genome editing in Plasmodium falciparum. Parasites & Vectors 9, 198-016-1487-4.Google Scholar
Ma, Y, Zhang, L and Huang, X (2014) Genome modification by CRISPR/Cas9. The FEBS Journal 281, 51865193.Google Scholar
Mogollon, CM, van Pul, FJ, Imai, T, Ramesar, J, Chevalley-Maurel, S, de Roo, GM, Veld, SA, Kroeze, H, Franke-Fayard, BM, Janse, CJ and Khan, SM (2016) Rapid generation of marker-free P. falciparum fluorescent reporter lines using modified CRISPR/Cas9 constructs and selection protocol. PLoS ONE 11, e0168362.Google Scholar
Nacer, A, Claes, A, Roberts, A, Scheidig-Benatar, C, Sakamoto, H, Ghorbal, M, Lopez-Rubio, JJ and Mattei, D (2015) Discovery of a novel and conserved Plasmodium falciparum exported protein that is important for adhesion of PfEMP1 at the surface of infected erythrocytes. Cellular Microbiology 17, 12051216.Google Scholar
Ng, CL, Siciliano, G, Lee, MC, de Almeida, MJ, Corey, VC, Bopp, SE, Bertuccini, L, Wittlin, S, Kasdin, RG, Le Bihan, A, Clozel, M, Winzeler, EA, Alano, P and Fidock, DA (2016) CRISPR-Cas9-modified pfmdr1 protects Plasmodium falciparum asexual blood stages and gametocytes against a class of piperazine-containing compounds but potentiates artemisinin-based combination therapy partner drugs. Molecular Microbiology 101, 381393.Google Scholar
Ng, H and Dean, N (2017) Dramatic improvement of CRISPR/Cas9 editing in Candida albicans by increased single guide RNA expression. mSphere 2, e0038516. eCollection 2017 Mar–Apr.Google Scholar
Nihongaki, Y, Kawano, F, Nakajima, T and Sato, M (2015) Photoactivatable CRISPR-Cas9 for optogenetic genome editing. Nature Biotechnology 33, 755760.Google Scholar
O'Connell, MR, Oakes, BL, Sternberg, SH, East-Seletsky, A, Kaplan, M and Doudna, JA (2014) Programmable RNA recognition and cleavage by CRISPR/Cas9. Nature 516, 263266.Google Scholar
Patterson, AG, Yevstigneyeva, MS and Fineran, PC (2017) Regulation of CRISPR-Cas adaptive immune systems. Current Opinion in Microbiology 37, 17.Google Scholar
Pawlowic, MC, Vinayak, S, Sateriale, A, Brooks, CF and Striepen, B (2017) Generating and maintaining transgenic cryptosporidium parvum parasites. Current Protocols in Microbiology 46, 20B.2.1-20B.2.32.Google Scholar
Peng, D, Kurup, SP, Yao, PY, Minning, TA and Tarleton, RL (2014) CRISPR-Cas9-mediated single-gene and gene family disruption in Trypanosoma cruzi. mBio 6, e0209714.Google Scholar
Pineda, M, Moghadam, F, Ebrahimkhani, MR and Kiani, S (2017) Engineered CRISPR systems for next generation gene therapies. ACS Synthetic Biology 6, 16141626.Google Scholar
Polstein, LR and Gersbach, CA (2015) A light-inducible CRISPR-Cas9 system for control of endogenous gene activation. Nature Chemical Biology 11, 198200.Google Scholar
Port, F and Bullock, SL (2016) Creating heritable mutations in drosophila with CRISPR-Cas9. Drosophila: Methods and Protocols. Methods in Molecular Biology 1478, 145160.Google Scholar
Qi, LS, Larson, MH, Gilbert, LA, Doudna, JA, Weissman, JS, Arkin, AP and Lim, WA (2013) Repurposing CRISPR as an RNA-guided platform for sequence-specific control of gene expression. Cell 152, 11731183.Google Scholar
Roos, DS, Donald, RG, Morrissette, NS and Moulton, AL (1994) Molecular tools for genetic dissection of the protozoan parasite Toxoplasma gondii. Methods in Cell Biology 45, 2763.Google Scholar
Roos, DS, Sullivan, WJ, Striepen, B, Bohne, W and Donald, RG (1997) Tagging genes and trapping promoters in Toxoplasma gondii by insertional mutagenesis. Methods 13, 112122.Google Scholar
Salsman, J and Dellaire, G (2017) Precision genome editing in the CRISPR era. Biochemistry and Cell Biology (Biochimie et biologie cellulaire) 95, 187201.Google Scholar
Semenova, E, Jore, MM, Datsenko, KA, Semenova, A, Westra, ER, Wanner, B, van der Oost, J, Brouns, SJ and Severinov, K (2011) Interference by clustered regularly interspaced short palindromic repeat (CRISPR) RNA is governed by a seed sequence. Proceedings of the National Academy of Sciences of the United States of America 108, 1009810103.Google Scholar
Sexton, T and Cavalli, G (2015) The role of chromosome domains in shaping the functional genome. Cell 160, 10491059.Google Scholar
Shen, B, Brown, KM, Lee, TD and Sibley, LD (2014 a) Efficient gene disruption in diverse strains of Toxoplasma gondii using CRISPR/CAS9. mBio 5, e0111414.Google Scholar
Shen, B, Buguliskis, JS, Lee, TD and Sibley, LD (2014 b) Functional analysis of rhomboid proteases during Toxoplasma invasion. mBio 5, e0179514.Google Scholar
Shen, B, Brown, K, Long, S and Sibley, LD (2017) Development of CRISPR/Cas9 for efficient genome editing in Toxoplasma gondii. Methods in Molecular Biology (Clifton, NJ) 1498, 79103.Google Scholar
Sidik, SM, Hackett, CG, Tran, F, Westwood, NJ and Lourido, S (2014) Efficient genome engineering of Toxoplasma gondii using CRISPR/Cas9. PLoS ONE 9, e100450.Google Scholar
Sidik, SM, Huet, D, Ganesan, SM, Huynh, MH, Wang, T, Nasamu, AS, Thiru, P, Saeij, JP, Carruthers, VB, Niles, JC and Lourido, S (2016) A genome-wide CRISPR screen in toxoplasma identifies essential apicomplexan genes. Cell 166, 14231435. e12.Google Scholar
Singer, M, Marshall, J, Heiss, K, Mair, GR, Grimm, D, Mueller, AK and Frischknecht, F (2015) Zinc finger nuclease-based double-strand breaks attenuate malaria parasites and reveal rare microhomology-mediated end joining. Genome Biology 16, 249-015-0811-1.Google Scholar
Slaymaker, IM, Gao, L, Zetsche, B, Scott, DA, Yan, WX and Zhang, F (2016) Rationally engineered Cas9 nucleases with improved specificity. Science (New York, NY) 351, 8488.Google Scholar
Soete, M, Hettman, C and Soldati, D (1999) The importance of reverse genetics in determining gene function in apicomplexan parasites. Parasitology 118(suppl.), S53S61.Google Scholar
Soldati, D and Boothroyd, JC (1993) Transient transfection and expression in the obligate intracellular parasite, Toxoplasma gondii. Science 260, 349351.Google Scholar
Straimer, J, Lee, MC, Lee, AH, Zeitler, B, Williams, AE, Pearl, JR, Zhang, L, Rebar, EJ, Gregory, PD, Llinas, M, Urnov, FD and Fidock, DA (2012) Site-specific genome editing in Plasmodium falciparum using engineered zinc-finger nucleases. Nature Methods 9, 993998.Google Scholar
Striepen, B (2013) Time to tackle cryptosporidiosis. Nature 503, 189191.Google Scholar
Suarez, CE, Bishop, RP, Alzan, HF, Poole, WA and Cooke, BM (2017) Advances in the application of genetic manipulation methods to apicomplexan parasites. International Journal for Parasitology 47, 701710.Google Scholar
Theodos, CM, Griffiths, JK, D'Onfro, J, Fairfield, A and Tzipori, S (1998) Efficacy of nitazoxanide against Cryptosporidium parvum in cell culture and in animal models. Antimicrobial Agents and Chemotherapy 42, 19591965.Google Scholar
Veiga, MI, Dhingra, SK, Henrich, PP, Straimer, J, Gnadig, N, Uhlemann, AC, Martin, RE, Lehane, AM and Fidock, DA (2016) Globally prevalent PfMDR1 mutations modulate Plasmodium falciparum susceptibility to artemisinin-based combination therapies. Nature Communications 7, 11553.Google Scholar
Vinayak, S, Pawlowic, MC, Sateriale, A, Brooks, CF, Studstill, CJ, Bar-Peled, Y, Cipriano, MJ and Striepen, B (2015) Genetic modification of the diarrhoeal pathogen Cryptosporidium parvum. Nature 523, 477480.Google Scholar
Volz, JC, Yap, A, Sisquella, X, Thompson, JK, Lim, NT, Whitehead, LW, Chen, L, Lampe, M, Tham, WH, Wilson, D, Nebl, T, Marapana, D, Triglia, T, Wong, W, Rogers, KL and Cowman, AF (2016) Essential role of the PfRh5/PfRipr/CyRPA complex during plasmodium falciparum invasion of erythrocytes. Cell Host & Microbe 20, 6071.Google Scholar
Wagner, JC, Platt, RJ, Goldfless, SJ, Zhang, F and Niles, JC (2014) Efficient CRISPR-Cas9-mediated genome editing in Plasmodium falciparum. Nature Methods 11, 915918.Google Scholar
Wang, J, Li, X, Zhao, Y, Li, J, Zhou, Q and Liu, Z (2015) Generation of cell-type-specific gene mutations by expressing the sgRNA of the CRISPR system from the RNA polymerase II promoters. Protein & Cell 6, 689692.Google Scholar
Wang, JL, Huang, SY, Behnke, MS, Chen, K, Shen, B and Zhu, XQ (2016 a) The past, present, and future of genetic manipulation in Toxoplasma gondii. Trends in Parasitology 32, 542553.Google Scholar
Wang, JL, Huang, SY, Li, TT, Chen, K, Ning, HR and Zhu, XQ (2016 b) Evaluation of the basic functions of six calcium-dependent protein kinases in Toxoplasma gondii using CRISPR-Cas9 system. Parasitology Research 115, 697702.Google Scholar
Wright, AV, Sternberg, SH, Taylor, DW, Staahl, BT, Bardales, JA, Kornfeld, JE and Doudna, JA (2015) Rational design of a split-Cas9 enzyme complex. Proceedings of the National Academy of Sciences of the United States of America 112, 29842989.Google Scholar
Wright, AV, Nunez, JK and Doudna, JA (2016) Biology and applications of CRISPR systems: harnessing nature's toolbox for genome engineering. Cell 164, 2944.Google Scholar
Xu, L, Zhao, L, Gao, Y, Xu, J and Han, R (2017) Empower multiplex cell and tissue-specific CRISPR-mediated gene manipulation with self-cleaving ribozymes and tRNA. Nucleic Acids Research 45, e28.Google Scholar
Yoshioka, S, Fujii, W, Ogawa, T, Sugiura, K and Naito, K (2015) Development of a mono-promoter-driven CRISPR/Cas9 system in mammalian cells. Scientific Reports 5, 18341.Google Scholar
Zetsche, B, Volz, SE and Zhang, F (2015) A split-Cas9 architecture for inducible genome editing and transcription modulation. Nature Biotechnology 33, 139142.Google Scholar
Zhang, C, Xiao, B, Jiang, Y, Zhao, Y, Li, Z, Gao, H, Ling, Y, Wei, J, Li, S, Lu, M, Su, XZ, Cui, H and Yuan, J (2014) Efficient editing of malaria parasite genome using the CRISPR/Cas9 system. mBio 5, e0141414.Google Scholar
Zhang, C, Gao, H, Yang, Z, Jiang, Y, Li, Z, Wang, X, Xiao, B, Su, XZ, Cui, H and Yuan, J (2017 a) CRISPR/cas9 mediated sequential editing of genes critical for ookinete motility in Plasmodium yoelii. Molecular and Biochemical Parasitology 212, 18.Google Scholar
Zhang, T, Gao, Y, Wang, R and Zhao, Y (2017 b) Production of guide RNAs in vitro and in vivo for CRISPR using ribozymes and RNA polymerase II promoters. Bio-protocol 7, e2148.Google Scholar
Zhang, YW, Kim, K, Ma, YF, Wittner, M, Tanowitz, HB and Weiss, LM (1999) Disruption of the Toxoplasma gondii bradyzoite-specific gene BAG1 decreases in vivo cyst formation. Molecular Microbiology 31, 691701.Google Scholar