• No results found

University of Groningen Opportunities and Challenges of Epigenetic Editing in Human Diseases Goubert, Désirée

N/A
N/A
Protected

Academic year: 2021

Share "University of Groningen Opportunities and Challenges of Epigenetic Editing in Human Diseases Goubert, Désirée"

Copied!
30
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Opportunities and Challenges of Epigenetic Editing in Human Diseases

Goubert, Désirée

DOI:

10.33612/diss.173201281

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from it. Please check the document version below.

Document Version

Publisher's PDF, also known as Version of record

Publication date: 2021

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Goubert, D. (2021). Opportunities and Challenges of Epigenetic Editing in Human Diseases: Towards the Curable Genome. University of Groningen. https://doi.org/10.33612/diss.173201281

Copyright

Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policy

If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum.

(2)

8

147043 Goubert BNW.indd 190

(3)

G

Geenneerraall DDiissccuussssiioonn aanndd FFuuttuurree PPeerrssppeeccttiivveess

General Discussion and

Future Perspectives

(4)

In many human diseases epigenetic regulation of genes goes wrong (1-4), which can have severe consequences on gene expression and cellular behaviour as well as on sensitivity to treatment and development of resistance (5, 6). The research presented in this thesis focusses on identifying and reversing epigenetic malfunctions that are associated with disturbed gene expression patterns in human diseases. The overall aim was to use Epigenetic Editing to reprogram specific target genes by rewriting epigenetic marks, without changing the primary DNA sequence (Chapter 1).

In Chapter 2 the scope of the EpiPredict consortium, of which this PhD project was part, is

described in detail. The overall focus within EpiPredict was to uncover the role of epigenetic regulation in the development of resistance to endocrine therapy in oestrogen receptor positive (ER+) breast cancer. A systems medicine approach was employed by combining multidisciplinary research strategies and next generation technologies (epigenetic/genetic profiling, protein pathway activation, metabolic pathway profiling, gene-specific epigenetic interference technologies and computational approaches). Results from these combined efforts include detailed procedures and protocols describing powerful tools (Single-Molecule RNA FISH, creation of dCas9-expressing cells lines, Chromatin Immunoprecipitation and High-Throughput Sequencing) which can be used to study epigenetic regulation (7-9). Profiling studies within the consortium provided us with valuable target genes, involved in the development of endocrine therapy resistance such as HES1, ESR1 (10), YY1 (6) and CD44 (11). Several sgRNAs were developed and tested for these genes (Figure 1). The sgRNAs designed to target ESR1 were based on previous studies describing successful targeting of Zinc Finger (ZF) proteins that were designed in our laboratory (12). ZF proteins are naturally occurring transcription factors which are, besides CRISPR-dCas9 sgRNAs, used as a programmable DNA targeting platform to guide effector domains to specific target genes (Chapter 3). Both the

artificial transcriptional repressor SKD (not possessing enzymatic activity but recruiting other epigenetics enzymes), as well as targeted methylation by variants of the DNA methyltransferase M.SssI were able to downregulate the expression of ESR1 in MCF-7 breast cancer cells, but not in ovarian cancer SKOV3 cells using our designed ZFs (12). The observed downregulation of ESR1 could not be repeated using sgRNAs together with CRISPR-dCas9 coupled to any artificial transcription factor (ATF) or to an epigenetic enzyme, reflecting the differences between the targeting platforms and the experimental set ups: The previously

(5)

used ZFs were stably expressed in the cells and thus were constantly present, whilst the sgRNAs were transiently transfected (to be able to assess long-term effects of the constructs, even when they are no longer expressed in the cells). Further, the relatively small size of ZF proteins (13, 14) compared to the larger size of CRISPR-dCas9 (approximately 8 times larger than ZFs) could explain the difference in e.g. gaining DNA accessibility. Issues of delivery, sgRNA efficiency and context-dependency are described in detail in this Chapter.

Due to lack of consistent repressive or inducible effects, or high variability of the effect on specifically the CD44 gene we did not continue to study HES1, ESR1, YY1 or CD44. Yet, the possibility to predict and avoid resistance to endocrine therapy will be a step closer to a so-called tailored therapy for breast cancer patients. Such epigenetic insights could be translated into diagnostic or prognostic tools to differentiate patients for their likelihood to develop endocrine resistance and predict the efficiency of additional drugs to counteract resistance in high risk patients.

To this end, the CD44 sgRNAs are currently further investigated in DKFZ, as this target is known to be associated with breast cancer invasiveness and stemness (15) and therefore might be important for endocrine resistance acquisition. Recent findings indicate an increase in tumour cells expressing high CD44 levels under oestrogen deprived conditions, which mimics a resistant environment (11). Early findings from DKFZ (16), under supervision of Prof. Dr. Stefan Wiemann, show that re-sensitization of resistant cells could be achieved using CRISPR tools. In long-term oestrogen deprived (LTED) MCF-7 cells (mimicking a resistant breast cancer phenotype) stably expressing a fusion of dCas9 and a histone methyltransferase (G9A), the targeting of CD44 with the sgRNAs designed in our lab resulted in repression of

CD44. The treated cells were more sensitive to oestrogen–deprived conditions, leading to a

significantly decreased proliferation compared to control cells which were transfected with a non-targeting control sgRNA. This finding strongly suggests the involvement of CD44 in the maintenance of endocrine therapy resistance in MCF-7 LTED cell lines.

(6)

Figure 9: Schematic representation of target gene promoter regions and locations of the respective sgRNAs (Green

arrows) and Zinc Finger Proteins (Blue arrows). CG islands (CGI) are represented by the red bar and the locations of the Transcription Start Site (TSS) is depicted.

(7)

We further focussed on using Epigenetic Editing in other target genes with known roles in cancer, as well as in fibrosis and airway diseases, including PLAU (Chapter 4), PLOD2 (Chapter 5), and UCHL1 (Chapter 6).

Achieving functional effects using Epigenetic Editing

Increased levels of PLAU have been associated with a worse prognosis and an increase in aggressiveness, metastasis, and invasion of breast cancer (17-22). By upregulating PLAU in non-invasive, hormone-sensitive, oestrogen receptor α positive luminal A MCF-7 breast cancer cells, using the artificial transcriptional activator dCas9-VP64, the oncogenic activity of this gene was indeed confirmed (Chapter 4). We observed an increased migration of breast

cancer cells with an increased PLAU expression, indicating a more aggressive phenotype. This is in line with other reports which describe induction of PLAU using untargeted approaches in other cell lines (23-30). These reports make use of PLAU activating compounds that also affect other genes which could lead to unwanted side effects in a clinical setting. In some of these reports, the increased migratory capacity after PLAU upregulation was successful in promoting wound healing (27, 30), indicating a potential application of PLAU induced upregulation. Our targeted PLAU-induction platform using dCas9-VP64 could thus be of use in pathological conditions such as ischemic brain injury (31-33), lung fibrosis (34), male infertility (35), type 2 diabetes mellitus (36) and diabetic keratopathy (37).

Next, we set out to repress PLAU in hormone-insensitive, triple negative MDA-MB-231 breast cancer cells to inhibit the migration potential of these aggressive cells. No significant downregulation by using the non-catalytic KRAB repressor (SKD), nor targeted DNA or histone methylation, using the DNA methyltransferase M.SssI or the histone 3 lysine 9 methyltransferase G9A, could be accomplished at this point. This is most likely because of poor delivery of our constructs. At the beginning of this project we were well aware of the low transfection efficiency in MDA-MB-231 cells, as is also shown by others (38-40). To overcome this problem, we created stable cells expressing dCas9- effector domains (EDs), a procedure that is described in detail in Chapter 7. Cell lines stably expressing dCas9-EDs are

expected to improve screening approaches, as dCas9 plasmids are relatively large making efficient delivery using transient transfections more difficult (8, 41). Several cell lines have

(8)

been transformed in our lab into stable cells continuously expressing several dCas9-EDs, however we were not able to show improved transient transfection efficiency compared to wild-type cells. For example, a more efficient repression of PLOD2 was observed in Chapter 5

using double transient transfection (plasmids expressing dCas9-SKD and sgRNAs) in wild-type HEK293T cells as compared to repression of PLOD2 in stable dCas9-SKD expressing HEK293T cells. Also stable expression of dCas9-VP64 in BEAS-2B cells in Chapter 6 did not significantly

improve on the UCHL1 induction achieved with transient transfection of this ED, despite the 100 times higher expression of dCas9 in stable cells compared to transiently transfected cells. Despite increasing the amount of sgRNA transfected in an attempt to improve upon our induced effects, we did not see a consistently higher gene induction or repression in any of the other genes tested in the stable cells. This effect could illustrate that the maximum effect of induction or repression of gene activity that can be achieved using transient transfection has already been reached in wild-type cells.

Since lentiviral transduction efficiency is known to be much higher than transient transfection (42-44), we set out to transduce MDA-MB-231 cells stably expressing dCas9-EDs a second time with lentiviral sgRNAs targeted to the PLAU promoter. No significant repression was observed, even though a trend towards downregulation was present in cells expressing the DNA methyltransferase Q147L or the histone 3 lysine 9 methyltransferase G9A. M.SssI-Q147L is a mutated variant of the CG-specific prokaryotic DNA methyltransferase M.SssI which has ~10% activity of the wild type M.SssI, the catalytically inactive mutant is M.SssI-E186A, which is used as a control. The lower DNA binding affinity of Q147L has been proven to increase the specificity of targeted DNA methylation over the wildtype M.SssI (45). In other cellular contexts and for other genes, these EDs did show repressive effects, which already shows some of the chromatin context-dependency that affects the efficiency of Epigenetic Editing, which is explained in more depth later in this Chapter. For example dCas9-G9A, but not it’s catalytic mutant, was able to achieve mitotically stable silencing of SPDEF in the type II alveolar carcinoma A549 cell lines as well as in MCF-7 cells (46) and M.SssI-Q147L, but not it’s catalytically inactive mutant (dCas9-E186A), induced gene repression of PLOD2 in fibroblasts, human embryonic kidney cells (HEK293T) and MDA-MB-231 breast cancer cells, but not in MCF-7 cells (Chapter 5).

(9)

In a second attempt to repress PLAU in MDA-MB-231 cells, we used Nucleofection® electroporation to transfect MDA-MB-231 dCas9-ED stable cells to increase the efficiency of the sgRNA delivery. This physical form of transfection employs an electrical force on the cells to create transient pores in the cell membrane, enhancing the uptake of molecular tools such as Epigenetic Editing constructs. Using this method, many promising in vivo gene editing applications are already being developed in mice (47-51). Because this method can be quite toxic for cells, it would be mainly suitable for ex vivo applications in humans, in which patient cells are taken out of the body, treated, and put back into the patient, as is shown by ongoing clinical trials using this physical form of transfection (52-55). Unfortunately, we did not see any effect on PLAU expression in any of the cell lines transfected by Nucleofection electroporation. In our Nucleofection experiments, we transfected three different fluorescently-tagged plasmids: a GFP (green fluorescence) -tagged control plasmid, a GFP-tagged sgRNA plasmid (empty, not targeted towards any gene) and a mCherry (red fluorescence) -tagged dCas9-NED (No Effector Domain). When cells successfully take up these plasmids, the fluorescent signals can be measured and efficiency of transfection can be determined. Although the transfection efficiency of cells Nucleofected with the small GFP control plasmid was sufficiently high (±60-70%), no fluorescent signal could be observed for the other two plasmids, despite using various concentrations of plasmid and different Nucleofection programs. Nucleofection of HEK293T with the same three plasmids showed similar results, with only very weak fluorescent signals being visible from the GFP-tagged sgRNA plasmid and the mCherry-tagged dCas9-NED plasmid. This suggests both plasmids were not properly transcribed within the cells. Future research should focus on optimizing delivery of targeted Epigenetic Editing tools to change gene expression and subsequently achieve changes in cell functioning.

Delivery of Epigenetic Editing constructs

One promising virus-free delivery method that could be further investigated for this purpose is the use of ligand-directed targeting approaches which target e.g. cancer cell specific surface receptors. These approaches, which might have therapeutic potential, make use of liposomes and other nanoparticles (56-58). Because of this, the therapeutic tools are more efficiently taken up by the cells, and are protected from degradation by the biological environment (56).

(10)

This method is not limited to only targeting cancer cells and holds the possibility to selectively target specific tissues (59, 60). These methods are attractive for clinical use as uptake by patients is improved compared to e.g. virus-based delivery methods (61). However, a possible disadvantage is the shorter in vivo life-time some of these extracellular vesicles can have compared to adeno-associated virus (AAV) -mediated delivery (60, 62). Also, for CRISPR-dCas9 based approaches such novel delivery techniques are showing promise for in vivo gene activation. Kretzmann et al. used dendritic polymers in a targeted intravenous approach to reactivate the tumour suppressor genes MASPIN and CCN6 in a sustained and efficient way in a mouse model of breast cancer (63). This reactivation led to an efficient and long-term cancer growth inhibition with negligible toxicity, outlining a targeted and effective method with potential to treat aggressive malignancies. Specifically in the difficult to transfect MDA-MB-231 cells an ultrasound targeted microbubble destruction (UTMD) technology has recently been optimized to deliver constructs in these cells (40). Here, Zhang et al. make use of the mechanical effects of ultrasound, combined with acoustically responsive microbubbles or droplets, creating transient pores in the cell membrane which enhances introduction of molecular tools. This technology has already shown great promise for the delivery of drugs or genes in models of solid tumours (64, 65), obesity (66) and cardiovascular disease (67). The most promising in vitro delivery system for Epigenetic Editing tools at the moment is adeno-associated viral (AAV) delivery, which is also the most commonly used vector for packaging and delivery of CRISPR components in vivo (68-72). For in vitro based research, lentiviral delivery is equally efficient, however upon infection lentiviruses randomly integrate into the genome of the host cell, leading to stable expression of the Epigenetic Editing tool. The adeno-associated virus is a non-pathogenic virus with a very mild immune response capable of delivering constructs in a tissue-specific way, with a high efficiency and without integration into the host genome, which could leads to harmful side-effects (73-75). Despite these wonderful properties, only very few Epigenetic Editing studies have successfully delivered CRISPR-dCas9-EDs in vivo using AAV delivery. This can be attributed to the limited packaging capacity of these AAV vectors (~5 kb) and the large size of the Epigenetic Editing constructs (76-78). Ongoing improvements such as smaller dCas9 orthologues and a split dCas9 system (in which the dCa9 enzyme is split, one part will be coupled to the ED, the other to the sgRNAs and the platform will only be active when both components bind) will

(11)

ultimately decide the success of AAV delivery of Epigenetic Editing constructs in a clinical setting (79).

Technical improvements

In view of the many possibilities Epigenetic Editing holds, technical improvements have been assessed in this thesis. For example in Chapter 6, we tested different experimental strategies

to modulate the gene expression of UCHL1, a gene with relevance in cancer and airway diseases. One such an improvement is the insertion of additional RNA hairpins into the sgRNA plasmid to form the sgRNA2.0 system (80). These RNA hairpins are recognised by RNA-specific binding proteins, which can be coupled to the same type of effector domains as to CRISPR-dCas9. The sgRNA2.0 is thus capable of recruiting both the RNA-specific binding protein effectors as well as the CRISPR-dCas9-ED, which can lead to amplification of the efficacy of induced effects (80-82). For example, Konermann et al. showed that targeting of sgRNA2.0 to

Neurog2 with a combination of both MS2-VP64 and dCas9-VP64 resulted in an additive effect,

leading to 12-fold higher increase compared to targeting only dCas9-VP64 (80). In our two cell lines however, MS2-targeting of p65-HSF1 (a transcriptional activator) did not improve the VP64-induction of UCHL1 gene expression even though the MS2-tagged p65-HSF1 on its own did induce gene expression up to 4-fold. This could indicate that either the maximum level of induction is reached in these cells, or and more likely, the lack of proper controls to ensure efficient delivery of the constructs (e.g. fluorescent tags to sort successfully transfected cells). To avoid the current difficulties of efficient delivery of large or multiple components, researchers can turn to methods of selection such as antibiotic resistance assays and fluorescence activated cell sorting (FACS) to enrich the transfected cell population (83). FACS can be performed upon delivery of fluorescent tags such as GFP or mCherry fused to the (d)Cas9 (84-86), to the sgRNAs (87), or to both allowing dual fluorescent tag sorting (88, 89). Cells that have successfully taken up the fluorescently-tagged constructs, will emit fluorescence at a specific wavelength when they are excited by a laser, after which they can be isolated from the bulk population and further investigated (90, 91). FACS, in our case, was the most successful improvement that we tested to increase upon the induction of UCHL1 using CRISPR-dCas9 in Chapter 6. The FACS sorting procedure utilizing indirect dCas9-mCherry

fusions resulted in an increased efficiency to measure gene expression changes in transfected

(12)

cells, displaying a 6- and 15-fold induced UCHL1 expression for the lung cancer H1299 and lung epithelial BEAS-2B cell lines, respectively, compared to the 2.2- and 2.8-fold increase observed in bulk cells, that were not sorted. In Chapter 4, no fluorescent signal could be

observed after nucleofection of the larger GFP-tagged sgRNA plasmid, or mCherry-tagged dCas9-NED, however this is likely because the constructs were not successfully delivered into MDA-MB-231 cells. Several fluorescently-tagged CRISPR-dCas9-ED have been created by us

(Chapters 4 and 6) that can be used to increase the population of successfully transfected

cells using FACS.

Sustainability of Epigenetic Editing

Besides delivery, another important challenge to overcome before epigenetic reprogramming can be developed into a straightforward technology for effective and specific interventions, is the potential dependency on genomic context to enable sustained transcription modulation. Sustained transcriptional modulation has been achieved in several in vivo models (71, 92, 93), for example, AAV delivery of dCas9-KRAB targeting Pcsk9 (a regulator of cholesterol levels) in liver cells of adult mice, resulted in reduction of Pcsk9 gene expression and reduction of cholesterol levels for a period of at least 24 weeks (71). Xenograft studies, in which NUDE mice were injected with tumour cells that stably express either a DNA methyltransferase or empty vector demonstrated stable SOX2 repression with maintenance of DNA methylation and long-term breast tumour growth inhibition, which lasted for more than 100 days after implantation of the tumour cells (92). Reactivation of Sim1 and Mc4r in haploinsufficient mice (in which loss-of-function mutations in one gene copy (as opposed to two) can lead to reduced amounts of protein and, consequently, disease) using AAV delivery of dCas9-VP64 into the hypothalamus of these mice led to reversal of the obesity phenotype normally seen in mice lacking expression of these genes, which was still observed nine months after injection of the constructs (93).

All the above examples, however, make use of the episomally maintained AAV or stably engineered cells, not reflecting true epigenetic reprogramming. The current consensus holds that to achieve sustained effects by epigenetic reprogramming, multiple effector domains (e.g. KRAB, DNA methyltransferase and/or histone modifiers) are required (75, 94-96). Transient expression of combinations of these effector domains resulted in synergistic effects

(13)

and long-term repression of several genes by deposition of repressive histone marks and de

novo DNA methylation. For clinical applications however, the requirement of multiple

components would be a serious limitation. In Chapter 5, we thus tested whether the fibrosis-

and cancer-associated gene PLOD2 can be repressed in a sustained manner by the individual effector domains DNA methyltransferase M.SssI, or the non-catalytic KRAB repressor (SKD). Both effector domains when fused to PLOD2-targeting ZFs, were able to repress PLOD2 also after 10 days of TGFβ1 stimulation (which induces PLOD2 expression in fibrosis (97-99)). To ensure the expression of the constructs in all analysed cells, the ZF-fusions were stably expressed by fibroblasts and MDA-MB-231 cells under the control of a doxycycline-inducible expression system. This entails that the expression of the ZFs is induced by adding doxycycline to the medium in which the cells grow. However, leakiness is a frequently observed phenomenon of this system. We indeed found similar levels of PLOD2 repression in cells with or without doxycycline, preventing us from making conclusions on the sustainability of SKD- or M.SssI-induced effects on PLOD2 expression using this system. We therefore investigated sustainability of the effects using the transient sgRNA expression system in HEK293T cells stably expressing dCas9-SKD or dCas9-M.SssI variants. Upon targeting PLOD2 in engineered HEK293T cells, SKD led to a sustained repression (after 2 and 12 days), while repression induced by M.SssI seems to take a longer time (only detectable after 12 days). This was a very striking observation, especially given that the current paradigm of SKD induced repression holds that transcriptional repressive effects are transient in somatic and cancer cells (75, 94, 96, 100-104). Using KRAB as an effector domain, Thakore et al., showed silencing of a cholesterol regulating gene (Pcsk9) in the liver of adult mice for a duration of at least 24 weeks upon delivery using AAV vectors (71). However, as mentioned before, these delivery vectors are episomally maintained, preventing conclusions from being drawn about the mitotic stability of the effect of KRAB itself. We therefore set out to confirm in a truly transient system, the sustainability of SKD-induced PLOD2 repression, by transfecting the plasmids to express dCas9-SKD as well as the sgRNAs into wildtype HEK293T cells. This resulted not only in a sustained repression, as previously seen in stable dCas9-SKD-expressing cells, the repression was also more efficient using double transient transfection (plasmids expressing dCas9-SKD and sgRNAs) as compared to repression of PLOD2 in stable dCas9-SKD expressing cells. Using the truly transient (double-transfection) CRISPR-dCas9 system, we showed that SKD as well as M.SssI targeting is sufficient to induce sustained PLOD2 repression. Targeted

(14)

transcriptional modulation using either M.SssI or SKD thus has the potential to evolve into anti-PLOD2 therapeutics against fibrosis.

Chromatin context-dependency of Epigenetic Editing

A critical hurdle to overcome before Epigenetic Editing can be implemented into the clinic is the gene and chromatin dependent optimization that is required, as targeting effects seem to be rather context-dependent. This context-dependency refers to several aspects including the local chromatin state and efficiency of sgRNAs (109, 110), the effect of the targeted ED or the edited epigenetic mark (111) and transcriptional state of the target gene (112). The chromatin state refers to the way DNA is modified and packed into the nucleus, and this influences the efficiency of DNA-targeting constructs, including CRISPR-dCas9, as well as their intended outcome (efficient localisation of EDs and successful deposition of epigenetic marks at the intended site does not always lead to the same functional effect) (113-116).

The chromatin context-dependency is already apparent in the first steps of the Epigenetic Editing approach i.e. the design of the DNA-binding platform to target genes of interest. For example, in Chapter 5, designing a different set of sgRNAs, which target the DNA at a

sequence similar to ZFs that were shown to successfully repress PLOD2 expression, did not lead to any repression in HEK293T cells stably expressing dCas9-SKD. Also in Chapter 6 we

could not improve the 5-fold induction in gene expression that was initially achieved by targeting dCas9-VP64 to UCHL1, by designing and combining sgRNAs that bind other genomic sequences. In MCF-7 cells (low PLAU expression), using the same PLAU targeting sgRNAs lead to significant induction of gene expression, whilst no significant repression could be achieved in MDA-MB-231 cells (high PLAU expression) (Chapter 4). As sgRNA design software does not

take the different chromatin contexts in different cell models into account, nor the type of effector domain that is targeted, the sgRNA design rules could be improved based on more systematic screening efforts (114, 117-119).

The effect of a targeted effector domain is also influenced by this context-dependency. For example, a combination of PRDM9 and DOT1L targeted to PLOD2 in C33A cervical cancer cells resulted in a significant and sustained upregulation (111). Using the same effector domains,

(15)

separately or in combination, did not lead to a short-term or long-term induction of PLAU in MCF-7 breast cancer cells (Chapter 4), whilst induction by dCas9-VP64 used in the same set

of experiments, as a positive control, did give a significant transient induction in MCF-7 cells. Transfection of only dCas9-SKD together with PLOD2 targeting sgRNAs gave a sustained repression of PLOD2 in HEK293T human embryonic kidney cells and MCF-7 cells, but no sustained repression was observed when SKD was targeted to another gene (SPDEF) in MCF-7 cells (Chapter 5). Co-transfection of dCas9-M.SssI-Q147L with dCas9-SKD, but not either of

these EDs individually, lead to a significant and sustained repression of UCHL1 in HEK293T, but not H1299 lung cancer cells (Chapter 6). Targeting methylated or unmethylated sites with

dCas9-VP64 has been shown to result in a difference in the efficiency of transcriptional upregulation of a gene (111). Hypermethylated genes ICAM1 and RASSF1a were not upregulated by targeting dCas9-VP64, whilst unmethylated genes EpCAM and PLOD2 did show a significant upregulation (111). This might suggest that dCas9-VP64 is not able to access the promoters of hypermethylated genes, explaining the lack of effect upon its targeting. Finally, the effect of epigenetic marks and whether or not they are maintained can also be chromatin context-dependent, so whilst an ED can lead to the deposition of epigenetic marks on a given target gene, the functional outcome can be different in e.g. different cell types. For example, efficient PLOD2 DNA methylation was detected in fibroblasts expressing either ZF7-M.SssI or ZF8-ZF7-M.SssI, even though at day 10 gene expression repressive effects of ZF8-ZF7-M.SssI were lost (Chapter 5). However, the difference in efficiency of induced DNA methylation

(63.4% induced by ZF7-M.SssI compared to 48.7% methylation induced by ZF8-M.SssI) could also explain the loss of PLOD2 repression seen in the latter. This could imply that to achieve sustained repression of the PLOD2 gene by targeting the M.SssI enzyme, a threshold amount of DNA methylation has to be induced and/or CpGs critical for transcription initiation need to be methylated. Another explanation could be that the target location of ZF7 was more responsive to deposited DNA methylation and subsequently lead to gene repression. Furthermore, in fibroblasts targeting of ZF7-M.SssI increased the presence of repressive histone marks H3K9me3 and H3K27me3, whilst this was not the case upon targeting ZF7-M.SssI in cancer cells (Chapter 5). This could be due to differences in growth kinetics of the

cells, which is much faster in the cancer cells compared to fibroblasts, or by the continuous TGFß1 stimulation applied to the fibroblasts but not to the cancer cells. Another example of context-dependent epigenetic marks is shown in Chapter 6. Here, an efficient UCHL1 DNA

(16)

methylation of 70% in HEK293T cells, using a combination of M.SssI-Q147L and dCas9-SKD, lead to a significant 40% gene repression in the sorted cell population. An efficient methylation of 60% in H1299 cells however, did not lead to a repressive effect, despite the fact that a region was targeted where DNA methylation is generally associated with UCHL1 repression (120) (Chapter 6). Similarly, sustained repression was achieved using targeted DNA

methylation in some studies (92, 121-123), but this could not be confirmed for other targeted genes (87, 124). These examples show that also the maintenance of epigenetic reprogramming is chromatin context-dependent.

Genetic vs. Epigenetic Editing

While, besides ex vivo, now also the first test of in-body (in vivo) gene editing in humans has shown encouraging results, one might be wondering what the added advantage of Epigenetic Editing is over the more developed application of gene editing. Early results of these in vivo gene editing trials suggest that the cells of the patients were successfully modified and that the that gene editing is safe in humans (125-127). Gene editing can correct errors in the genome by changing the DNA sequence, whilst Epigenetic Editing changes the expression of genes, possibly with sustained effects, without changing the DNA sequence. Ongoing clinical trials can be registered on https://clinicaltrials.gov/, where they get an unique identification or NTC number, and as of 2020 a handful of gene editing trials have entered clinical trials. For example, researchers from Sangamo Therapeutics are using gene editing techniques to insert a healthy copy of the iduronate-2-sulfatase gene into liver cells, and in this way treat people with Hunter syndrome a condition in which patients are not able to break down complex sugars (NCT03041324) (128, 129). At the same time, clinical trials are also ongoing to treat hemophilia B (NCT02695160) and Hurler syndrome (NCT02702115). In all these three in vivo cases, ZFs are targeted toward the disease-causing gene and delivered into the patients via virus based approaches. For clinical applications, despite a difficult start, viral vectors (Adeno Associated Viruses and Lentiviruses) are reaching widespread acceptance (130). Recently, other researchers have also tested CRISPR-based gene editing inside a person’s body for the first time. Allergan and Editas Medicine treated the first patient with Leber congenital amaurosis 10, an inherited type of blindness caused by mutations in the CEP290 gene (NCT03655678). Microscopic droplets carrying an inactivated virus are administered by a

(17)

subretinal injection in one eye. Results are not yet known, and the primary goal of this early-stage trial is to test the safety of the approach, which is partly why only one eye is treated initially (131).

Despite these impressive and very important steps forward, gene editing techniques also pose several important downsides that could limit their clinical potential. Given the fact that DNA alterations are permanent, one of the most critical considerations is the specificity of gene editing (132). There is also an uncertainty about the potentially harmful effects on the offspring, especially when off-target effects occur in the germline (133). Another potential issue is that genome editing by CRISPR–Cas9 induces a p53-mediated immune response and cell cycle arrest (134). These and other concerns with regard to gene editing were even more ignited after the birth of the first gene-edited babies in China, when He Jiankui acted against official regulations (released by China’s health and science ministries in 2003) as announced at the Second International Summit on Human Genome Editing in Hong Kong on 29 November 2018 (135). Gene editing has thus been the subject of many ethical debates and concerns worldwide. Unfortunately, rules and legislation about the use and implementation of this technique are lagging behind.

These downsides of the available gene editing techniques stress the need for alternative therapeutic approaches. Sustained correction of aberrant expression levels of disease-associated genes as well as reversal of the downstream effects of a genetic mutation can be achieved through Epigenetic Editing. Extensive elaboration of the concept as well as applications of Epigenetic Editing are discussed throughout this thesis. Through Epigenetic Editing a natural, endogenous expression control is exerted by epigenetic rewriting of the gene’s own promoters/enhancers, resembling more physiological conditions in normal cells (136). Further, multiple isoforms of a gene can be targeted, which could be essential in achieving the desired functional outcome (137). Even though Epigenetic Editing suffers from the same downside with off-target effects as genetic editing, its reversible nature allows for an easier correction or reversal of potential side-effects. Unwanted edited epigenetic marks can be removed or replaced again in a targeted and gene-specific manner, without long-lasting effects on the gene or DNA sequence. Despite this reversible nature, edited epigenetic marks can be inherited by the next cell generation, and stably maintained throughout cell

(18)

divisions (71, 75, 111), resulting in durable and long-lasting modification of gene expression. This is different from epigenetic marks being inherited by the offspring, which is only possible through alterations within the germ cells (the egg and sperm). This is highly controversial and is not, nor should it be, allowed in humans. Being inherited by the next cell generation however, is important for clinical purposes where the ideal therapy would be to administer Epigenetic Editing constructs in a one-and-done approach, where patients are treated once, after which the modified epigenetic marks are remembered and maintained on the target gene. In this way, the final result of Epigenetic Editing on gene expression can be similar to gene editing approaches with regard to permanent reprogramming of gene expression, but now without changing the DNA sequence.

The ability to permanently reprogram gene expression is especially apparent during mammalian development. Even though most cells within an organism contain the same DNA, there are many different cell types, making up the various tissues and organs. Epigenetic modifications underlie cell identity by switching cell-type specific genes on or off during cell divisions (e.g. proliferating liver cells, remain liver cells) (138). Because of this intrinsic mechanism of maintenance of epigenetic marks, also edited epigenetic marks can remain stable on the DNA or histone tails, even after removal of the Epigenetic Editing tool. Furthermore, the extend of the rewritten epigenetic marks can spread along the target gene (139), as endogenous epigenetic enzymes are recruited to the target side (140, 141), also contributing to the maintenance of the rewritten epigenetic environment. Taken these considerations into account, even though it has not yet evolved into a clinical translation such as gene editing, Epigenetic Editing is ethically considered less invasive (safer) as no genetic changes are introduced. Reports on preclinical therapeutic effects upon viral delivery of epigenetic editors together with an improved understanding of biology, are facilitating in vivo Epigenetic Editing and transcriptional modulation (79). For example, Fragile X syndrome (FXS), the most common genetic form of intellectual disability in males and often associated with social anxiety and autism, is caused by silencing of the FMR1 gene by DNA hypermethylation. Removing of these methyl groups by lentiviral delivery of dCas9-Tet1 (targeted demethylation) restores the expression of FMR1 in multiple FXS patient derived induced pluripotent stem cell (iPSC) lines. After maturation of these stem cells into neurons (by well-established differentiation protocols) and transplanting them into the brains of new-born

(19)

mice, the matured neurons continued to express FMR1 three months later (142). The same approach could have promise for treating Rett syndrome, which generally is lethal in boys and related with autism in girls, and results from a loss-of-function mutation in one copy of the X chromosome-linked gene MECP2. Re-expressing specific regions in the normally silenced X chromosome could turn on the intact version of MECP2, without affecting other X chromosome genes, to treat Rett syndrome symptoms (143). Another in vivo example of preclinical effects using Epigenetic Editing is provided by Zeisberg et al., who show that re-expression of hypermethylated genes Rasal1 and Klotho, using lentiviral delivered dCas9-Tet3, attenuates disease progression in a mouse model of kidney fibrosis (144). Hypermethylation of these genes has also been associated with fibrosis in heart (145) and liver (146) as well as progression of various forms of cancer (147), demonstrating that CRISPR-dCas9-based gene-specific DNA demethylation has a broad application spectrum and may be useful to combat these other diseases as well. In neurodegenerative disorders like e.g. Alzheimer’s disease, targeted upregulation of Dlg4 has been shown to lead to functional differences like improved learning in aged and Alzheimer's disease mice (136). Injection of herpes simplex virus (HSV) particles containing ZF-VP64 constructs targeted to the Dlg4 gene into the hippocampus of these mice, improved their performance in learning-related tasks and rescued memory deficits. These, and other (45, 63, 71-73, 92, 93, 148-153), examples show the potential of using Epigenetic Editing and targeted transcriptional modulation as a therapy to treat human diseases (Table 1).

(20)

Table 1: Noncomprehensive summary showing the therapeutic potential of using Epigenetic Editing in vivo to treat

human diseases

Human Disease Purpose Platform + Targeting

ED Delivery method Result Animal Reference

Breast Cancer Targeted DNA methylation of SOX2 ZF-DNMT3A

Xenograft transplantation of

tumour cells that stably express the

constructs

Long-lasting oncogenic repression and DNA

methylation Adult Mice Stolzenburg et al. 2015 (92) Several human diseases

and biological processes

Develop strategy for targeted DNA

demethylation dCas9-TET1 Lentiviral

Activation of a methylation reporter by demethylation of

its promoter

Adult

Mice 2016 (148) Liu et al.

Cancer and repression ofTranscriptional activation cancer-related genes

dCas9 dCas9-VP64

Injection of cells that stably

express the constructs

Down-regulation of Trp53: accelerating disease onset, reducing survival, resistance to

treatment + Increased expression of

MGMT: resistance to

treatment and shorter survival Adult Mice

Braun et al. 2016. (149)

Alzheimer’s disease Transcriptional activation of Dlg4 ZF-VP64 Herpes Simplex Viral

Upregulation of Dlg4 and improved performance in learning-and memory-related

tasks

Adult

Mice Bustos et al. 2017 (136) Several human diseases,

including cancer

Develop strategy for locus-specific, rapid and efficient

DNA methylation dCas9- M.SssI-Q147L Zygote micro-injection of plasmids

Targeted DNA methylation at specific CpGs of the imprinted

locus Igf2/H19

Mice

embryo’s Lei et al. 2017 (4) Several human diseases

including type I diabetes, acute kidney

injury, and muscular dystrophy

Transcriptional activation

of endogenous genes MS2/P65/HSF1 Adeno-Associated Viral

Rescue levels of gene expression (e.g. Klotho) +

Compensate for genetic defects (Utrophin) + Alter cell

fates (Pdx1)

Adult

Mice Liao et al. 2017 (62) Fragile X syndrome Re-expression of the silenced FMR1 gene dCas9-Tet1 Lentiviral Targeted demethylation and re-expression of FMR1 Adult Mice Liu et al. 2018

(142) Kidney fibrosis hypermethylated genes Re-expression of

Rasal1 and Klotho dCas9-Tet3 Lentiviral

Gene specific demethylation and re-expression + attenuation of fibrosis

Adult

Mice Xu et al. 2018 (144) Retinitis Pigmentosa Targeted repression of Nrl dCas9-KRAB Adeno-Associated Viral the retina + reprogramming of In situ Nrl gene repression in

rods into cone-like cells

Adult

Mice al. 2018 (72) Moreno et

Liver injury Transcriptional activation of hepatocyte-

specific genes in the liver dCas9-VP64

Adeno-Associated Viral in dCas9 Mice

Activation of endogenous gene expression, development of

hepatocellular carcinomas when oncogenes were

activated Adult Mice Wangensteen et al. 2018 (150) Investigation of gene function in the mammalian brain Targeted repression of

genes in neurons dCas9-KRAB

Lentiviral, stereotactic

injections

Inactivation of genes fundamental for neurotransmitter release with

high efficiency

Adult

Mice al. 2018 (152) Zheng et

Obesity Transcriptional activation of Sim1 and Mc4r dCas9-VP64 Adeno-Associated Viral

Increased expression of Sim1 and Mc4r + Reversal of the

obesity phenotype in haploinsufficient mice

Adult

Mice al. 2019 (93) Matharu et

Neuropsychiatric diseases, such as addiction, depression, schizophrenia, and Alzheimer’s disease Transcriptional activation of targeted genes in neurons dCas9-VP64 Lentiviral, stereotactic injections

Specific and large-scale control of gene expression profiles within the central nervous

system

Adult

Mice Savell et al. 2019 (151) Breast Cancer Transcriptional activation of tumour suppressor

genes dCas9-VP64 Dendritic Polymers

Reactivation of MASPIN and

CCN6

+ cancer growth inhibition

Adult

(21)

Concluding Remarks

In order to propel Epigenetic Editing towards clinically addressing human pathologies, the challenges discussed in this thesis should be taken into account. Overcoming critical challenges such as chromatin context-dependency, delivery and specificity will be essential in the coming years and at the same time public debate about the applications and ethical considerations should be encouraged. Science is not finished until it is communicated, and only through outreach can science truly make an impact. As researchers, we all have the obligation and responsibility to act in an ethically responsible way, to educate the future generation and to engage with society. Bridging the gaps in research, communication and legislation will ultimately decide the success of “the curable genome” using Epigenetic Editing in human diseases.

References

1. Egger G, Liang G, Aparicio A, Jones PA. Epigenetics in human disease and prospects for epigenetic therapy. Nature. 2004;429(6990):457-63.

2. Portela A, Esteller M. Epigenetic modifications and human disease. Nat Biotechnol. 2010;28(10):1057-68.

3. Dekker AD, De Deyn PP, Rots MG. Epigenetics: the neglected key to minimize learning and memory deficits in Down syndrome. Neurosci Biobehav Rev. 2014;45:72-84.

4. Keating ST, El-Osta A. Epigenetics and metabolism. Circ Res. 2015;116(4):715-36.

5. Badia E, Oliva J, Balaguer P, Cavaillès V. Tamoxifen resistance and epigenetic modifications in breast cancer cell lines. Curr Med Chem. 2007;14(28):3035-45.

6. Patten DK, Corleone G, Győrffy B, Perone Y, Slaven N, Barozzi I, et al. Enhancer mapping uncovers phenotypic heterogeneity and evolution in patients with luminal breast cancer. Nat Med. 2018;24(9):1469-80.

7. Beckman W, Vuist IM, Kempe H, Verschure PJ. Cell-to-Cell Transcription Variability as Measured by Single-Molecule RNA FISH to Detect Epigenetic State Switching. Methods Mol Biol. 2018;1767:385-93.

8. Goubert D, Koncz M, Kiss A, Rots MG. Establishment of Cell Lines Stably Expressing dCas9-Fusions to Address Kinetics of Epigenetic Editing. Methods Mol Biol. 2018;1767:395-415. 9. Patten DK, Corleone G, Magnani L. Chromatin Immunoprecipitation and High-Throughput

Sequencing (ChIP-Seq): Tips and Tricks Regarding the Laboratory Protocol and Initial Downstream Data Analysis. Methods Mol Biol. 2018;1767:271-88.

10. Magnani L, Frige G, Gadaleta RM, Corleone G, Fabris S, Kempe MH, et al. Acquired CYP19A1 amplification is an early specific mechanism of aromatase inhibitor resistance in ERα metastatic breast cancer. Nat Genet. 2017;49(3):444-50.

11. Hong SP, Chan TE, Lombardo Y, Corleone G, Rotmensz N, Bravaccini S, et al. Single-cell transcriptomics reveals multi-step adaptations to endocrine therapy. Nat Commun. 2019;10(1):3840.

12. Falahi F, Hospers GAP, Rots MG. Epigenetic Editing of estrogen receptor-alpha gene in breast cancer as an innovative tool to modulate

its expression level. Groningen, The Netherlands: The university of Groningen; 2014.

(22)

13. Falahi F, Sgro A, Blancafort P. Epigenome engineering in cancer: fairytale or a realistic path to the clinic? Front Oncol. 2015;5:22.

14. Mussolino C, Morbitzer R, Lütge F, Dannemann N, Lahaye T, Cathomen T. A novel TALE nuclease scaffold enables high genome editing activity in combination with low toxicity. Nucleic Acids Res. 2011;39(21):9283-93.

15. Orian-Rousseau V. CD44 Acts as a Signaling Platform Controlling Tumor Progression and Metastasis. Front Immunol. 2015;6:154.

16. Sofyali E. GLYATL1 promotes

endocrine therapy resistance in breast cancer. Germany: the Ruperto Carola University of Heidelberg; 2019.

17. McMahon BJ, Kwaan HC. Components of the Plasminogen-Plasmin System as Biologic Markers for Cancer. Adv Exp Med Biol. 2015;867:145-56.

18. Mazar AP, Ahn RW, O'Halloran TV. Development of novel therapeutics targeting the urokinase plasminogen activator receptor (uPAR) and their translation toward the clinic. Curr Pharm Des. 2011;17(19):1970-8.

19. Mahmood N, Mihalcioiu C, Rabbani SA. Multifaceted Role of the Urokinase-Type Plasminogen Activator (uPA) and Its Receptor (uPAR): Diagnostic, Prognostic, and Therapeutic Applications. Front Oncol. 2018;8:24.

20. Moquet-Torcy G, Tolza C, Piechaczyk M, Jariel-Encontre I. Transcriptional complexity and roles of Fra-1/AP-1 at the uPA/Plau locus in aggressive breast cancer. Nucleic Acids Res. 2014;42(17):11011-24.

21. Duffy MJ, Duggan C, Mulcahy HE, McDermott EW, O'Higgins NJ. Urokinase plasminogen activator: a prognostic marker in breast cancer including patients with axillary node-negative disease. Clin Chem. 1998;44(6 Pt 1):1177-83.

22. Duffy MJ, O'Grady P, Devaney D, O'Siorain L, Fennelly JJ, Lijnen HJ. Urokinase-plasminogen activator, a marker for aggressive breast carcinomas. Preliminary report. Cancer. 1988;62(3):531-3.

23. Li T, Jiang S. Effect of bFGF on invasion of ovarian cancer cells through the regulation of Ets-1 and urokinase-type plasminogen activator. Pharm Biol. 2010;48(2):161-5.

24. Kim MJ, Kim DH, Na HK, Surh YJ. TNF-alpha induces expression of urokinase-type plasminogen activator and beta-catenin activation through generation of ROS in human breast epithelial cells. Biochem Pharmacol. 2010;80(12):2092-100.

25. Pulukuri SM, Gorantla B, Dasari VR, Gondi CS, Rao JS. Epigenetic upregulation of urokinase plasminogen activator promotes the tropism of mesenchymal stem cells for tumor cells. Mol Cancer Res. 2010;8(8):1074-83.

26. Madhyastha R, Madhyastha H, Nakajima Y, Omura S, Maruyama M. Curcumin facilitates fibrinolysis and cellular migration during wound healing by modulating urokinase plasminogen activator expression. Pathophysiol Haemost Thromb. 2010;37(2-4):59-66.

27. Wang L, Zhang Y, Wang W, Zhu Y, Chen Y, Tian B. Gemcitabine treatment induces endoplasmic reticular (ER) stress and subsequently upregulates urokinase plasminogen activator (uPA) to block mitochondrial-dependent apoptosis in Panc-1 cancer stem-like cells (CSCs). PLoS One. 2017;12(8):e0184110.

28. Chen LC, Shibu MA, Liu CJ, Han CK, Ju DT, Chen PY, et al. ERK1/2 mediates the lipopolysaccharide-induced upregulation of FGF-2, uPA, MMP-2, MMP-9 and cellular migration in cardiac fibroblasts. Chem Biol Interact. 2019;306:62-9.

29. Sawai H, Okada Y, Funahashi H, Matsuo Y, Takahashi H, Takeyama H, et al. Interleukin-1alpha enhances the aggressive behavior of pancreatic cancer cells by regulating the alpha6beta1-integrin and urokinase plasminogen activator receptor expression. BMC Cell Biol. 2006;7:8. 30. Radha KS, Madhyastha HK, Nakajima Y, Omura S, Maruyama M. Emodin upregulates urokinase

plasminogen activator, plasminogen activator inhibitor-1 and promotes wound healing in human fibroblasts. Vascul Pharmacol. 2008;48(4-6):184-90.

(23)

31. Cho E, Lee KJ, Seo JW, Byun CJ, Chung SJ, Suh DC, et al. Neuroprotection by urokinase plasminogen activator in the hippocampus. Neurobiol Dis. 2012;46(1):215-24.

32. Diaz A, Merino P, Manrique LG, Ospina JP, Cheng L, Wu F, et al. A Cross Talk between Neuronal Urokinase-type Plasminogen Activator (uPA) and Astrocytic uPA Receptor (uPAR) Promotes Astrocytic Activation and Synaptic Recovery in the Ischemic Brain. J Neurosci. 2017;37(43):10310-22.

33. Merino P, Yepes M. Urokinase-type Plasminogen Activator Induces Neurorepair in the Ischemic Brain. J Neurol Exp Neurosci. 2018;4(2):24-9.

34. Horowitz JC, Tschumperlin DJ, Kim KK, Osterholzer JJ, Subbotina N, Ajayi IO, et al. Urokinase Plasminogen Activator Overexpression Reverses Established Lung Fibrosis. Thromb Haemost. 2019;119(12):1968-80.

35. Zhao K, Liu Y, Xiong Z, Hu L, Xiong CL. Tissue-specific inhibition of urokinase-type plasminogen activator expression in the testes of mice by inducible lentiviral RNA interference causes male infertility. Reprod Fertil Dev. 2017;29(11):2149-56.

36. Wu CZ, Ou SH, Chang LC, Lin YF, Pei D, Chen JS. Deficiency of Urokinase Plasminogen Activator May Impair beta Cells Regeneration and Insulin Secretion in Type 2 Diabetes Mellitus. Molecules. 2019;24(23).

37. Sun H, Mi X, Gao N, Yan C, Yu FS. Hyperglycemia-suppressed expression of Serpine1 contributes to delayed epithelial wound healing in diabetic mouse corneas. Invest Ophthalmol Vis Sci. 2015;56(5):3383-92.

38. Köster F, Schröder A, Finas D, Hauser C, Diedrich K, Felberbaum R. Cell-specific enhancement of liposomal transfection by steroids in steroid receptor expressing cells. Int J Mol Med. 2006;18(6):1201-5.

39. Wang T, Larcher LM, Ma L, Veedu RN. Systematic Screening of Commonly Used Commercial Transfection Reagents towards Efficient Transfection of Single-Stranded Oligonucleotides. Molecules. 2018;23(10).

40. Zhang H, Li Y, Rao F, Liufu C, Wang Y, Chen Z. A novel UTMD system facilitating nucleic acid delivery into MDA-MB-231 cells. Biosci Rep. 2020;40(2).

41. Kroll C, Rathert P. Stable Expression of Epigenome Editors via Viral Delivery and Genomic Integration. Methods Mol Biol. 2018;1767:215-25.

42. Cao F, Xie X, Gollan T, Zhao L, Narsinh K, Lee RJ, et al. Comparison of gene-transfer efficiency in human embryonic stem cells. Mol Imaging Biol. 2010;12(1):15-24.

43. Zecchin D, Di Nicolantonio F. Transfection and DNA-mediated gene transfer. Methods Mol Biol. 2011;731:435-50.

44. Elegheert J, Behiels E, Bishop B, Scott S, Woolley RE, Griffiths SC, et al. Lentiviral transduction of mammalian cells for fast, scalable and high-level production of soluble and membrane proteins. Nat Protoc. 2018;13(12):2991-3017.

45. Lei Y, Zhang X, Su J, Jeong M, Gundry MC, Huang YH, et al. Targeted DNA methylation in vivo using an engineered dCas9-MQ1 fusion protein. Nat Commun. 2017;8:16026.

46. Song J, Cano Rodriguez D, Winkle M, Gjaltema RA, Goubert D, Jurkowski TP, et al. Targeted epigenetic editing of SPDEF reduces mucus production in lung epithelial cells. Am J Physiol Lung Cell Mol Physiol. 2016:ajplung.00059.2016.

47. Lambricht L, Lopes A, Kos S, Sersa G, Préat V, Vandermeulen G. Clinical potential of electroporation for gene therapy and DNA vaccine delivery. Expert Opin Drug Deliv. 2016;13(2):295-310.

48. Latella MC, Di Salvo MT, Cocchiarella F, Benati D, Grisendi G, Comitato A, et al. In vivo Editing of the Human Mutant Rhodopsin Gene by Electroporation of Plasmid-based CRISPR/Cas9 in the Mouse Retina. Mol Ther Nucleic Acids. 2016;5(11):e389.

49. Wu W, Lu Z, Li F, Wang W, Qian N, Duan J, et al. Efficient in vivo gene editing using ribonucleoproteins in skin stem cells of recessive dystrophic epidermolysis bullosa mouse model. Proc Natl Acad Sci U S A. 2017;114(7):1660-5.

(24)

50. Qin W, Wang H. Delivery of CRISPR-Cas9 into Mouse Zygotes by Electroporation. Methods Mol Biol. 2019;1874:179-90.

51. Hashimoto M, Takemoto T. Electroporation enables the efficient mRNA delivery into the mouse zygotes and facilitates CRISPR/Cas9-based genome editing. Sci Rep. 2015;5:11315. 52. Kebriaei P, Singh H, Huls MH, Figliola MJ, Bassett R, Olivares S, et al. Phase I trials using

Sleeping Beauty to generate CD19-specific CAR T cells. J Clin Invest. 2016;126(9):3363-76. 53. Xu L, Wang J, Liu Y, Xie L, Su B, Mou D, et al. CRISPR-Edited Stem Cells in a Patient with HIV

and Acute Lymphocytic Leukemia. N Engl J Med. 2019;381(13):1240-7.

54. Stadtmauer EA, Fraietta JA, Davis MM, Cohen AD, Weber KL, Lancaster E, et al. CRISPR-engineered T cells in patients with refractory cancer. Science. 2020;367(6481).

55. Fajrial AK, He QQ, Wirusanti NI, Slansky JE, Ding X. A review of emerging physical transfection methods for CRISPR/Cas9-mediated gene editing. Theranostics. 2020;10(12):5532-49. 56. Wang M, Glass ZA, Xu Q. Non-viral delivery of genome-editing nucleases for gene therapy.

Gene Ther. 2017;24(3):144-50.

57. Saleh AF, Lázaro-Ibáñez E, Forsgard MA, Shatnyeva O, Osteikoetxea X, Karlsson F, et al. Extracellular vesicles induce minimal hepatotoxicity and immunogenicity. Nanoscale. 2019;11(14):6990-7001.

58. Conway A, Mendel M, Kim K, McGovern K, Boyko A, Zhang L, et al. Non-viral Delivery of Zinc Finger Nuclease mRNA Enables Highly Efficient In Vivo Genome Editing of Multiple Therapeutic Gene Targets. Mol Ther. 2019;27(4):866-77.

59. Cheng Q, Wei T, Farbiak L, Johnson LT, Dilliard SA, Siegwart DJ. Selective organ targeting (SORT) nanoparticles for tissue-specific mRNA delivery and CRISPR-Cas gene editing. Nat Nanotechnol. 2020.

60. Göran Ronquist K. Extracellular vesicles and energy metabolism. Clin Chim Acta. 2019;488:116-21.

61. Ma XX, Xu JL, Jia YY, Zhang YX, Wang W, Li C, et al. Enhance transgene responses through improving cellular uptake and intracellular trafficking by bio-inspired non-viral vectors. J Nanobiotechnology. 2020;18(1):26.

62. van Gestel MA, Boender AJ, de Vrind VA, Garner KM, Luijendijk MC, Adan RA. Recombinant adeno-associated virus: efficient transduction of the rat VMH and clearance from blood. PLoS One. 2014;9(5):e97639.

63. Kretzmann JA, Evans CW, Moses C, Sorolla A, Kretzmann AL, Wang E, et al. Tumour suppression by targeted intravenous non-viral CRISPRa using dendritic polymers. Chem Sci. 2019;10(33):7718-27.

64. Fan CH, Chang EL, Ting CY, Lin YC, Liao EC, Huang CY, et al. Folate-conjugated gene-carrying microbubbles with focused ultrasound for concurrent blood-brain barrier opening and local gene delivery. Biomaterials. 2016;106:46-57.

65. Lin L, Fan Y, Gao F, Jin L, Li D, Sun W, et al. UTMD-Promoted Co-Delivery of Gemcitabine and miR-21 Inhibitor by Dendrimer-Entrapped Gold Nanoparticles for Pancreatic Cancer Therapy. Theranostics. 2018;8(7):1923-39.

66. Chen S, Bastarrachea RA, Shen JS, Laviada-Nagel A, Rodriguez-Ayala E, Nava-Gonzalez EJ, et al. Ectopic BAT mUCP-1 overexpression in SKM by delivering a BMP7/PRDM16/PGC-1a gene cocktail or single PRMD16 using non-viral UTMD gene therapy. Gene Ther. 2018;25(7):497-509.

67. Deng Q, Hu B, Cao S, Song HN, Chen JL, Zhou Q. Improving the efficacy of therapeutic angiogenesis by UTMD-mediated Ang-1 gene delivery to the infarcted myocardium. Int J Mol Med. 2015;36(2):335-44.

68. Senís E, Fatouros C, Große S, Wiedtke E, Niopek D, Mueller AK, et al. CRISPR/Cas9-mediated genome engineering: an adeno-associated viral (AAV) vector toolbox. Biotechnol J. 2014;9(11):1402-12.

(25)

69. Swiech L, Heidenreich M, Banerjee A, Habib N, Li Y, Trombetta J, et al. In vivo interrogation of gene function in the mammalian brain using CRISPR-Cas9. Nat Biotechnol. 2015;33(1):102-6. 70. Yang Y, Wang L, Bell P, McMenamin D, He Z, White J, et al. A dual AAV system enables the

Cas9-mediated correction of a metabolic liver disease in newborn mice. Nat Biotechnol. 2016;34(3):334-8.

71. Thakore PI, Kwon JB, Nelson CE, Rouse DC, Gemberling MP, Oliver ML, et al. RNA-guided transcriptional silencing in vivo with S. aureus CRISPR-Cas9 repressors. Nat Commun. 2018;9(1):1674.

72. Moreno AM, Fu X, Zhu J, Katrekar D, Shih YV, Marlett J, et al. In Situ Gene Therapy via AAV-CRISPR-Cas9-Mediated Targeted Gene Regulation. Mol Ther. 2018;26(7):1818-27.

73. Liao HK, Hatanaka F, Araoka T, Reddy P, Wu MZ, Sui Y, et al. In Vivo Target Gene Activation via CRISPR/Cas9-Mediated Trans-epigenetic Modulation. Cell. 2017;171(7):1495-507.e15. 74. Naso MF, Tomkowicz B, Perry WL, Strohl WR. Adeno-Associated Virus (AAV) as a Vector for

Gene Therapy. BioDrugs. 2017;31(4):317-34.

75. Amabile A, Migliara A, Capasso P, Biffi M, Cittaro D, Naldini L, et al. Inheritable Silencing of Endogenous Genes by Hit-and-Run Targeted Epigenetic Editing. Cell. 2016;167(1):219-32 e14. 76. Wu Z, Yang H, Colosi P. Effect of genome size on AAV vector packaging. Mol Ther.

2010;18(1):80-6.

77. Grieger JC, Samulski RJ. Packaging capacity of adeno-associated virus serotypes: impact of larger genomes on infectivity and postentry steps. J Virol. 2005;79(15):9933-44.

78. Colella P, Ronzitti G, Mingozzi F. Emerging Issues in AAV-Mediated. Mol Ther Methods Clin Dev. 2018;8:87-104.

79. Lau C-H, Suh Y. In vivo epigenome editing and transcriptional modulation using CRISPR technology. Transgenic research. 2018:1-21.

80. Konermann S, Brigham MD, Trevino AE, Joung J, Abudayyeh OO, Barcena C, et al. Genome-scale transcriptional activation by an engineered CRISPR-Cas9 complex. Nature. 2015;517(7536):583-8.

81. Zalatan JG, Lee ME, Almeida R, Gilbert LA, Whitehead EH, La Russa M, et al. Engineering complex synthetic transcriptional programs with CRISPR RNA scaffolds. Cell. 2015;160(1-2):339-50.

82. Xu X, Tao Y, Gao X, Zhang L, Li X, Zou W, et al. A CRISPR-based approach for targeted DNA demethylation. Cell Discov. 2016;2:16009.

83. Ran FA, Hsu PD, Wright J, Agarwala V, Scott DA, Zhang F. Genome engineering using the CRISPR-Cas9 system. Nat Protoc. 2013;8(11):2281-308.

84. Chen B, Gilbert LA, Cimini BA, Schnitzbauer J, Zhang W, Li GW, et al. Dynamic imaging of genomic loci in living human cells by an optimized CRISPR/Cas system. Cell. 2013;155(7):1479-91.

85. Anton T, Bultmann S, Leonhardt H, Markaki Y. Visualization of specific DNA sequences in living mouse embryonic stem cells with a programmable fluorescent CRISPR/Cas system. Nucleus. 2014;5(2):163-72.

86. Anton T, Leonhardt H, Markaki Y. Visualization of Genomic Loci in Living Cells with a Fluorescent CRISPR/Cas9 System. Methods Mol Biol. 2016;1411:407-17.

87. McDonald JI, Celik H, Rois LE, Fishberger G, Fowler T, Rees R, et al. Reprogrammable CRISPR/Cas9-based system for inducing site-specific DNA methylation. Biol Open. 2016;5(6):866-74.

88. Shao S, Zhang W, Hu H, Xue B, Qin J, Sun C, et al. Long-term dual-color tracking of genomic loci by modified sgRNAs of the CRISPR/Cas9 system. Nucleic Acids Res. 2016;44(9):e86. 89. Fu Y, Rocha PP, Luo VM, Raviram R, Deng Y, Mazzoni EO, et al. CRISPR-dCas9 and sgRNA

scaffolds enable dual-colour live imaging of satellite sequences and repeat-enriched individual loci. Nat Commun. 2016;7:11707.

90. Herzenberg LA, Sweet RG. Fluorescence-activated cell sorting. Sci Am. 1976;234(3):108-17.

(26)

91. Herzenberg LA, Parks D, Sahaf B, Perez O, Roederer M. The history and future of the fluorescence activated cell sorter and flow cytometry: a view from Stanford. Clin Chem. 2002;48(10):1819-27.

92. Stolzenburg S, Beltran AS, Swift-Scanlan T, Rivenbark AG, Rashwan R, Blancafort P. Stable oncogenic silencing in vivo by programmable and targeted de novo DNA methylation in breast cancer. Oncogene. 2015;34(43):5427-35.

93. Matharu N, Rattanasopha S, Tamura S, Maliskova L, Wang Y, Bernard A, et al. CRISPR-mediated activation of a promoter or enhancer rescues obesity caused by haploinsufficiency. Science. 2019;363(6424).

94. O'Geen H, Ren C, Nicolet CM, Perez AA, Halmai J, Le VM, et al. dCas9-based epigenome editing suggests acquisition of histone methylation is not sufficient for target gene repression. Nucleic Acids Res. 2017;45(17):9901-16.

95. Mlambo T, Nitsch S, Hildenbeutel M, Romito M, Müller M, Bossen C, et al. Designer epigenome modifiers enable robust and sustained gene silencing in clinically relevant human cells. Nucleic Acids Res. 2018;46(9):4456-68.

96. O'Geen H, Bates SL, Carter SS, Nisson KA, Halmai J, Fink KD, et al. Ezh2-dCas9 and KRAB-dCas9 enable engineering of epigenetic memory in a context-dependent manner. Epigenetics Chromatin. 2019;12(1):26.

97. van der Slot AJ, van Dura EA, de Wit EC, De Groot J, Huizinga TW, Bank RA, et al. Elevated formation of pyridinoline cross-links by profibrotic cytokines is associated with enhanced lysyl hydroxylase 2b levels. Biochim Biophys Acta. 2005;1741(1-2):95-102.

98. Remst DF, Blaney Davidson EN, Vitters EL, Blom AB, Stoop R, Snabel JM, et al. Osteoarthritis-related fibrosis is associated with both elevated pyridinoline cross-link formation and lysyl hydroxylase 2b expression. Osteoarthritis Cartilage. 2013;21(1):157-64.

99. Gjaltema RA, de Rond S, Rots MG, Bank RA. Procollagen Lysyl Hydroxylase 2 Expression Is Regulated by an Alternative Downstream Transforming Growth Factor β-1 Activation Mechanism. J Biol Chem. 2015;290(47):28465-76.

100. de Groote ML, Verschure PJ, Rots MG. Epigenetic Editing: targeted rewriting of epigenetic marks to modulate expression of selected target genes. Nucleic Acids Res. 2012;40(21):10596-613.

101. Groner AC, Tschopp P, Challet L, Dietrich JE, Verp S, Offner S, et al. The Krüppel-associated box repressor domain can induce reversible heterochromatization of a mouse locus in vivo. J Biol Chem. 2012;287(30):25361-9.

102. Stolzenburg S, Rots MG, Beltran AS, Rivenbark AG, Yuan X, Qian H, et al. Targeted silencing of the oncogenic transcription factor SOX2 in breast cancer. Nucleic Acids Res. 2012;40(14):6725-40.

103. Thakore PI, Black JB, Hilton IB, Gersbach CA. Editing the epigenome: technologies for programmable transcription and epigenetic modulation. Nat Methods. 2016;13(2):127-37. 104. Ecco G, Cassano M, Kauzlaric A, Duc J, Coluccio A, Offner S, et al. Transposable Elements and

Their KRAB-ZFP Controllers Regulate Gene Expression in Adult Tissues. Dev Cell. 2016;36(6):611-23.

105. Zuurmond AM, van der Slot-Verhoeven AJ, van Dura EA, De Groot J, Bank RA. Minoxidil exerts different inhibitory effects on gene expression of lysyl hydroxylase 1, 2, and 3: implications for collagen cross-linking and treatment of fibrosis. Matrix Biol. 2005;24(4):261-70.

106. Gilkes DM, Bajpai S, Wong CC, Chaturvedi P, Hubbi ME, Wirtz D, et al. Procollagen lysyl hydroxylase 2 is essential for hypoxia-induced breast cancer metastasis. Mol Cancer Res. 2013;11(5):456-66.

107. Chen Y, Terajima M, Yang Y, Sun L, Ahn YH, Pankova D, et al. Lysyl hydroxylase 2 induces a collagen cross-link switch in tumor stroma. J Clin Invest. 2015;125(3):1147-62.

108. Piersma B, Bank RA. Collagen cross-linking mediated by lysyl hydroxylase 2: an enzymatic battlefield to combat fibrosis. Essays Biochem. 2019;63(3):377-87.

(27)

109. Uusi-Mäkelä MIE, Barker HR, Bäuerlein CA, Häkkinen T, Nykter M, Rämet M. Chromatin accessibility is associated with CRISPR-Cas9 efficiency in the zebrafish (Danio rerio). PLoS One. 2018;13(4):e0196238.

110. Chen Y, Zeng S, Hu R, Wang X, Huang W, Liu J, et al. Using local chromatin structure to improve CRISPR/Cas9 efficiency in zebrafish. PLoS One. 2017;12(8):e0182528.

111. Cano-Rodriguez D, Gjaltema RA, Jilderda LJ, Jellema P, Dokter-Fokkens J, Ruiters MH, et al. Writing of H3K4Me3 overcomes epigenetic silencing in a sustained but context-dependent manner. Nat Commun. 2016;7:12284.

112. Fujita T, Yuno M, Fujii H. Allele-specific locus binding and genome editing by CRISPR at the p16INK4a locus. Sci Rep. 2016;6:30485.

113. Chen H, Kazemier HG, de Groote ML, Ruiters MH, Xu GL, Rots MG. Induced DNA demethylation by targeting Ten-Eleven Translocation 2 to the human ICAM-1 promoter. Nucleic Acids Res. 2014;42(3):1563-74.

114. Verkuijl SA, Rots MG. The influence of eukaryotic chromatin state on CRISPR-Cas9 editing efficiencies. Curr Opin Biotechnol. 2019;55:68-73.

115. Huisman C, van der Wijst MG, Schokker M, Blancafort P, Terpstra MM, Kok K, et al. Re-expression of Selected Epigenetically Silenced Candidate Tumor Suppressor Genes in Cervical Cancer by TET2-directed Demethylation. Mol Ther. 2016;24(3):536-47.

116. Ambrosio R, Damiano V, Sibilio A, De Stefano MA, Avvedimento VE, Salvatore D, et al. Epigenetic control of type 2 and 3 deiodinases in myogenesis: role of Lysine-specific Demethylase enzyme and FoxO3. Nucleic Acids Res. 2013;41(6):3551-62.

117. Radzisheuskaya A, Shlyueva D, Müller I, Helin K. Optimizing sgRNA position markedly improves the efficiency of CRISPR/dCas9-mediated transcriptional repression. Nucleic Acids Res. 2016;44(18):e141.

118. Xue L, Tang B, Chen W, Luo J. Prediction of CRISPR sgRNA Activity Using a Deep Convolutional Neural Network. J Chem Inf Model. 2019;59(1):615-24.

119. Menon AV, Sohn JI, Nam JW. CGD: Comprehensive guide designer for CRISPR-Cas systems. Comput Struct Biotechnol J. 2020;18:814-20.

120. Jiang W, Liu N, Chen XZ, Sun Y, Li B, Ren XY, et al. Genome-Wide Identification of a Methylation Gene Panel as a Prognostic Biomarker in Nasopharyngeal Carcinoma. Mol Cancer Ther. 2015;14(12):2864-73.

121. Rivenbark AG, Stolzenburg S, Beltran AS, Yuan X, Rots MG, Strahl BD, et al. Epigenetic reprogramming of cancer cells via targeted DNA methylation. Epigenetics. 2012;7(4):350-60. 122. Bintu L, Yong J, Antebi YE, McCue K, Kazuki Y, Uno N, et al. Dynamics of epigenetic regulation

at the single-cell level. Science. 2016;351(6274):720-4.

123. Saunderson EA, Stepper P, Gomm JJ, Hoa L, Morgan A, Allen MD, et al. Hit-and-run epigenetic editing prevents senescence entry in primary breast cells from healthy donors. Nat Commun. 2017;8(1):1450.

124. Kungulovski G, Nunna S, Thomas M, Zanger UM, Reinhardt R, Jeltsch A. Targeted epigenome editing of an endogenous locus with chromatin modifiers is not stably maintained. Epigenetics Chromatin. 2015;8:12.

125. Harmatz P. EMPOWERS: A phase 1/2 clinical trial of SB-318 ZFNmediated in vivo human genome editing for treatment of MPS I (Hurler syndrome). Orlando, FL: World Symposium 2019; 2019.

126. Muenzer J. CHAMPIONS: A phase 1/2 clinical trial with dose escalation of SB-913 ZFN-mediated in vivo human genome editing for treatment of MPS II (Hunter syndrome). Orlando, FL: World Symposium 2019; 2019.

127. Konkle BA, Stine K, Visweshwar N, Harrington T, Leavitt AD, Arkin S, et al. Initial results of the Alta study, a phase 1/2, open label, adaptive, dose-ranging study to assess the safety and tolerability of SB-525 gene therapy in adult subjects with hemophilia A. Melbourne, Australia: ISTH2019; 2019.

Referenties

GERELATEERDE DOCUMENTEN

In this study, we were able to silence SPDEF expression in the human alveolar epithelial cell line A549, using a novel strategy: engineered SPDEF targeting ZF proteins

(c) Relati- ve mRNA expression of ICAM1, RASSF1a, EpCAM in HEK293T and A549 cells and (d) PLOD2 in C33a cells, by the indicated dCas9 fusion protein co-transfected with a

However, modulation of the target gene by induction of histone marks in the chromatin context was demonstrated first in 2002, by inducing H3K9 methylation and causing gene

First, we reviewed in chapter 2 the most recent advances in targeting epigenetic effector domains to different regions in the genome, in order to alter gene expression..

In 2012, after finishing his master, he got the great opportunity to obtain a PhD position at the University Medical Center Groningen, from the Rijksuniversiteit Gro- ningen,

Opportunities and Challenges of Epigenetic Editing in Human Diseases Goubert,

In this consortium, 12 PhD students from eight different countries and scientific institutes have studied the role of epigenetic regulation in resistance development for

We will determine cellular heterogeneity and sub-cell type epigenetic state switching [29] and use CRISPR-dCas9-based Epigenetic Editing [30] to locally overwrite epigenetic