Abstract
Genomic loci associated with common traits and diseases are typically non-coding and likely impact gene expression, sometimes coinciding with rare loss-of-function variants in the target gene. However, our understanding of how gradual changes in gene dosage affect molecular, cellular, and organismal traits is currently limited. To address this gap, we induced gradual changes in gene expression of four genes using CRISPR activation and inactivation. Downstream transcriptional consequences of dosage modulation of three master trans-regulators associated with blood cell traits (GFI1B, NFE2, and MYB) were examined using targeted single-cell multimodal sequencing. We showed that guide tiling around the TSS is the most effective way to modulate cis gene expression across a wide range of fold-changes, with further effects from chromatin accessibility and histone marks that differ between the inhibition and activation systems. Our single-cell data allowed us to precisely detect subtle to large gene expression changes in dozens of trans genes, revealing that many responses to dosage changes of these three TFs are non-linear, including non-monotonic behaviours, even when constraining the fold-changes of the master regulators to a copy number gain or loss. We found that the dosage properties are linked to gene constraint and that some of these non-linear responses are enriched for disease and GWAS genes. Overall, our study provides a straightforward and scalable method to precisely modulate gene expression and gain insights into its downstream consequences at high resolution.
Introduction
Precision control of gene expression levels plays a pivotal role in defining cell type specificity and coordinating responses to external stimuli. Imbalances in this intricate regulation can underlie the genetic basis of both common and rare human disease. The vast majority of genetic variants associated with complex disease, as revealed by genome-wide association studies (GWAS), are located in noncoding regions, with likely gene regulatory effects 1. Previous studies have attempted to elucidate these effects by mapping genetic associations to gene expression 2,3, and more recently, CRISPR-based perturbations of GWAS loci have provided insights into their functional consequences 4. A major driver of rare genetic diseases are loss-of-function variants affecting one or both copies of the gene, leading to disease via dramatic reduction of functional gene dosage 5. The substantial overlap 6,7 and potential joint effects 8,9 of rare and common variants indicate a general link between different degrees of perturbation of gene dosage and disease phenotypes.
However, our understanding of the quantitative relationship between gradual changes of gene dosage and downstream phenotypes remains elusive for most human genes. Practical applications of the compelling allelic series concept to identify genes where increasingly deleterious mutations have increasing phenotypic effects have been limited by the sparsity of segregating variants with an impact on a given gene in the human population 10. Experimental characterization of gene function in model systems has predominantly relied on gene knock-out or knock-down approaches 11. While these studies have proven useful to identify dosage sensitive genes involved in cellular functions and disease 12–16, these approaches only provide a limited discrete relationship between the number of functional gene copies and a certain phenotype (eg. loss-of-function consequence vs. wild-type). However, such relationships are in fact determined by continuous dosage-to-phenotypes functions that, as suggested by a small number of previous experimental studies 17–19, can be complex and thus are challenging to infer from loss-/gain-of-function data.
Recently, new methods have enabled the gradual modulation of gene dosage in model systems 18,20–22, while large-scale insights into the downstream effects of dosage modulation have largely come from yeast 17 and bacteria 19,23, demonstrating that non-linear relationships between gene dosage and phenotype are common. In humans, the relationship between dosage and downstream phenotype is largely unexplored. Only a few limited studies 17–19 have dissected these consequences, for instance on the disease-associated transcription factor SOX2 24. Such work showed a non-linear relationship between dosage and multiple tiers of phenotypes, including DNA accessibility, RNA expression of downstream targets, rending the question if such phenomenon occurs with other transcription factors. More recently, similar evidence has been shown in the case of the NKX2-1 lineage factor with oncogenic role in lung adenocarcinoma 25. Generally, transcription factors represent a particularly compelling target for characterization of gene dosage effects. They are key regulators of cellular functions, enriched for disease associations 26 and often classified as haploinsufficient 27. Additionally, their effects can be measured by transcriptome analysis. However, our knowledge of their dosage-dependent effects on regulatory networks still remains limited, particularly regarding subtle dosage variation within their natural range 22.
In this study, we developed and characterised a scalable novel sgRNA design approach for gradually decreasing and increasing gene dosage with the CRISPRI inhibition (CRISPRi) and activation (CRISPRa) systems. We applied this to four genes, with single cell RNA-sequencing (scRNA-seq) as a cellular readout of downstream effects. While classic Perturb-Seq analyses have focused on gene knock-down effects, we assess the effects of gradual up- and downregulation of target genes. We uncovered a quantitative landscape of how gradual changes in transcription dosage lead to linear and non-linear response in downstream genes, including those associated with rare and complex disease, with potential effects on cellular phenotypes.
Results
Precise modulation and quantification of gene dosage using CRISPR and targeted multimodal single-cell sequencing
We selected four genes for gradual modulation of their dosage in the human erythroid progenitor cell line K562 28: GFI1B, NFE2, MYB and TET2. Two of the genes, GFI1B and NFE2, have been implicated in blood diseases and traits 29–31, and in our earlier work we identified a broad transcriptional response to inhibition of GWAS-overlapping enhancers to these genes 4. MYB is a key transcription factor 32 and a downstream target of GFI1B 4.TET2 has a role in DNA demethylation and is unrelated to these transcriptional networks and is included in this study as control with minimal expected trans effects. We refer to these four genes, targeted in cis for modulation of their regulation, as cis genes (Figure 1A).
To modulate the gene expression of the cis genes we use K562 cells expressing CRISPRi (KRAB-dCas9-MECP2) and CRISPRa (dCas9-VPR) systems (see Methods), both cell lines hashed with DNA conjugated antibodies against different surface proteins that allow pooled experiments. To obtain a wide range of dosage effects we used four different single guide RNA (sgRNA) design strategies (Figure 1B): 1) targeting the transcription start site (TSS) as in the standard CRISPRi/CRISPRa approach, 2) tiling sgRNAs +/− 1000 bp from the TSS, 3) targeting known cis-regulatory elements (CREs), and 4) using attenuated guides that target the TSS but contain mismatches to modulate their activity 18. We further included 5 non-targeting control (NTC) sgRNAs as negative controls.
The library of altogether 96 guides was transduced to a pool of K562-CRISPRi and K562-CRISPRa cells at low multiplicity of infection (MOI). After eight days, we performed ECCITE-seq (see Methods) to capture three modalities: cDNA, sgRNAs and surface protein hashes (oligo-tagged antibodies with unique barcodes against ubiquitously expressed surface proteins). Instead of sequencing the full transcriptome, we used target hybridization to capture a smaller fraction of the cDNA and obtain more accurate expression readouts at a feasible cost. The subset of selected transcripts were picked from the transcriptional downstream regulatory networks of GFI1B and NFE2 identified previously 4, maintaining similar patterns of co-expression correlation across co-expression clusters (see Methods, Figure S1A). We targeted a total of 94 transcripts (Figure 1A), including the four cis genes, 86 genes that represent trans targets of GFI1B and/or NFE2 4 (Figure S1A), LXH3 that is not expressed in blood progenitors, GAPDH that is highly expressed and often considered an invariable housekeeping gene and the dCas9-VPR or KRAB-dCas9-MeCP2 transcripts.
We used the protein hashes and the dCas9 cDNA (the presence or absence of the KRAB domain) to demultiplex and determine the cell line—CRISPRi or CRISPRa—cells containing a single sgRNA per cell were determined using a Gausian mixed model (see Methods). We applied standard QC approaches to the scRNA-seq data and demonstrated the success of the target capture (see Methods, Figure S1C). The final data set had 20,001 cells (10,647 CRISPRi and 9,354 CRISPRa), with an average of 81 and 86 cells with a unique sgRNA for the CRISPRa and CRISPRi, respectively (Figure S1D).
Gradual modulation of gene expression across a broad range with CRISPRi/a
Next, we calculated the expression fold change for each of the four cis genes targeted by each sgRNA in the two cell lines (CRISPRi/a), comparing each group of cells with its respective NTC sgRNA group (see Methods). We first confirmed that the sgRNAs targeting the transcription start site (TSS) up- and downregulated their targets (Figure 1C, Figure S1F). When looking at all sgRNAs at once, across the four genes, we observed a 2.3 fold range (Figure 1E), with minimum 72% reduction and maximum 174% increased expression (log2(FC) values from −1.83 to 0.80). However, the range varied between the genes, with GFI1B covering the widest range of gene expression changes (gene expression ranging between 0.28 to 1.42 fold), while MYB expression could not be pushed higher than 1.13 fold (Figure 1E). The direction of the effects were consistent with the cell lines of origin, where 98.88% of the significant perturbations (Wilcoxon rank test at 10% FDR, n = 89) were correctly predicted based on the direction of the target gene fold change. The predicted on- and off-target properties of the guides 33–35 did not correlate with the fold changes in the cis genes (Figure S2A), suggesting that the observed effects represent true cis-regulatory changes. The fold changes were also robust to the number of cells containing a particular sgRNA (Figure S2B, top).
We verified that the fold change estimation was not biassed depending on the expression level of the target gene at the single-cell level, which can vary due to drop-out effects or binary on/off effects of the KRAB-based CRISPRi system 20. By splitting cells with the same sgRNA based on the normalised expression of the cis gene (0 vs. >0 normalised UMIs, Figure S3A), we observed highly concordant transcriptome gene expression effects between the two groups (Figure S3B). This indicates that the dosage changes per guide were not primarily driven by the changing frequency of binary on/off effects, and the use of pseudo-bulk fold changes provides a robust estimation of cis gene fold changes. These patterns are further supported by the cells forming a gradient rather than distinct clusters on a UMAP (Figure S3C).
The fold change patterns differed between sgRNA designs (Figure 1D, left). As expected, sgRNAs targeting the TSS showed strong perturbations in gene expression. However, sgRNAs tiled +/− 1kb from the TSS provided a broader and more gradual range of up- and downregulation across the target genes, sometimes surpassing the effects of TSS-targeting sgRNAs. Attenuated sgRNAs with mismatch mutations resulted in a range of gene silencing effects in the CRISPRi line, as expected based on their original design 18. However, these attenuated sgRNAs did not exhibit such a dynamic range in the CRISPRa modality, although a significant correlation existed between the silencing or activating effect size and the distance of the mismatch from the protospacer adjacent motif (PAM) when considering all data points together (Figure S2C). The sgRNAs targeting distal cis-regulatory elements (CREs) showed both inhibiting and activating effects, even though both the CRISPRi and CRISPRa constructs were initially designed to inhibit or activate transcription from the promoter and initial gene body region. Nonetheless, the number of known CREs per gene is typically limited. Given its simplicity and the ability to achieve both up- and downregulation of the target gene, we consider the tiling sgRNA approach, with a simple design that only requires annotation of the TSS, as the best unbiased method for gradually modulating gene dosage with CRISPRi/a systems.
Cis determinants of dosage
Having designed guides targeting both distal and local neighbouring regulatory regions of the four transcription factors (TFs) and ensuring minimal bias in fold-changes due to sgRNA’s biochemical properties, we investigated the cis features that determine the strength of dosage perturbation. We observed substantial differences in the effects of the same guide on the CRISPRi and CRISPRa backgrounds, with no significant correlation between cis gene fold-changes (Figure 2A). However, in both modalities, the strongest effects on gene expression were observed when the guides were close to the transcription start site (TSS) (Figure 2B, excluding NTC and attenuated sgRNAs), although the peaks of strongest activation or repression differed between the modalities. In the CRISPRi modality, the maximum effect was located within the gene body at +238 bp from the TSS (Figure 2B, bottom), consistent with previous studies that used essentiality as a proxy for expression 36. However, in the CRISPRa modality, the maximum average fold changes occurred closer to the TSS at around −99 bp (Figure 2B, bottom), as also shown for CD45 37.
Enhancer, tiling and TSS sgRNAs targeted regions of the genome with different compositions of histone marks in K562 annotated by ENCODE 38 (Figure 2C). This allowed us to investigate the impact of chromatin state on the strength of cis gene dosage modulation. The magnitude of cis gene fold changes varied significantly depending on the presence of specific marks or peaks, which again differed between the two modalities (Figure 2D). In the CRISPRa cell line, the strongest effects were observed when guides were located in regions with open chromatin marks such as DNase or ATAC peaks. In contrast, the strongest repression by CRISPRi occurred in genomic regions with the presence of H3K27ac, H3K4me3, and H3K9ac marks. These differences may be explained by the distinct mechanisms of action of activator and repressor domains. MeCP2 and KRAB repressor domains recruit additional repressors that silence gene expression through chromatin remodelling activities such as histone deacetylation 39. On the other hand, the VPR activation fusion domain may only require Cas9 to scan the open chromatin and recruit RNA polymerase and additional transcription factors to activate transcription. Overall, while a few sgRNAs have a strong effect in both CRISPRi and CRISPRa cell lines, a single guide library containing guides optimised for both modalities enables a range of gradual dosage regulation. However, larger data sets are needed for more careful modelling of the ideal dosage modulation designs and to understand how both cis-regulatory features, feedback loops and other mechanisms contribute to the outcomes.
Trans responses of transcription factor dosage modulation
We then turned our attention to the remaining 91 genes captured by our custom panel and determined the relative expression fold change of each trans gene, compared to NTC in each unique guide perturbation (see Methods). Principal component analysis (PCA) performed on all pseudo-bulk fold changes demonstrated the removal of batch effects from the cell lines and revealed a clear direction of the cis gene dosage effect in the first three principal components (Figure S4B). This finding suggests that dosage modulation is the primary determinant of trans effects. The PCA indicated that the dosage modulation of GFI1B and MYB is reflected in opposite directions in PC1 and PC2, while the trans responses of NFE2 are captured by PC3.
Using a false discovery rate (FDR) cutoff of 0.05, all 91 trans genes except the neural specific TF LHX3 (negative control) exhibited a significant change in expression upon perturbation of any of the TFs. The observed trans-effects were well correlated with perturbations of these genes in other data sets (Figure S4C,D). Among all measured fold changes, the most extreme negative effect sizes were observed in cis genes, with the top 10 being predominantly reductions in GFI1B expression. This indicates that cis downregulation tended to surpass the endogenous expression limits. In contrast, the largest increases in gene expression were observed through trans mechanisms, where KLK1 and TUBB1 reached the largest expression values when GFI1B was strongly upregulated, or SPI1 and DAPK1 when GFI1B was strongly downregulated. These findings suggest that the CRISPRa approach did not reach a biological ceiling of overexpression.
Inspecting trans responses as a function of cis gene modulation, we observed that the number of expressed genes and the mean absolute expression changes of trans genes exhibited correlations gene-specific correlations with cis-gene dosage (Figure 3A, Figure S4E). Perturbations in GFI1B led to the most pronounced trans responses, with positive dosage changes resulting in larger effect sizes compared to decreasing TF gene expression, where the effect plateaued. NFE2 exhibited similar patterns but with smaller magnitude. In the case of MYB, trans responses were observed when decreasing the expression of this TF, but the effects of upregulation are largely unknown as we were unable to increase MYB expression beyond 0.35. As expected, given the unrelatedness of TET2 to the trans network, dosage modulation of this gene had minimal trans effects with the least pronounced trend when compared to TET2 dosage, so we excluded it from subsequent analyses.
Widespread non-linear dosage responses in trans regulatory networks
Upon clustering the changes in expression of trans genes based on the cis gene dosage change linked to each sgRNA, we identified distinct clusters exhibiting different dosage-response patterns (Figure 3B for GFI1B, Figure S5-8A for all cis genes). Further examination of the gene expression fold changes for each individual transgene in relation to the TF fold changes revealed a diverse range of response patterns (Figure 3C, Figure S5-8B for all cis genes). These responses exhibited both linear and nonlinear forms, including some instances of non-monotonic gene expression responses for certain trans genes within the GFI1B trans network (e.g., GATA2 in Figure 3C, Figure S9E).
To accurately characterise the dosage response, we employed both linear and nonlinear modelling approaches (Figure 3D), which allowed us to quantitatively assess the extent of nonlinear responses by comparing the goodness of fit of these models using the Akaike Information Criterion (AIC). For the nonlinear model, we utilised a sigmoid function with four free parameters (Figure 3D, right). These parameters represented the slope at the inflection point (b, indicating the rate of increase or decrease in expression), the minimum and maximum asymptotes (c and d, representing the lower and upper limits of fold change), and the value of cis gene expression at which the inflection point occurs (a). To prevent overfitting, we implemented a 10-fold cross-validation scheme, which yielded reliable predictions on the left-out data (pearson r = 0.71 to 0.88 for all trans genes in the GFI1B, MYB, and NFE2 networks, Figure S9C). Additionally, the predicted parameter a was centred around zero, as expected since the input data represents relative fold changes (Figure S10). Since a sigmoid function cannot capture non-monotonic responses, we employed a loess regression as an alternative approach for the few genes that exhibited non-monotonic responses (see Methods, Figure S9D,E). For the vast majority of genes, the sigmoid (or loess) fit was remarkably good, partially due to the low level of noise in the targeted scRNA-seq data.
We compared the performance of the linear vs. nonlinear models with the ΔAIC (AlClinear - AlCnonlinear), where positive ΔAIC means that the sigmoid model captures better the variance in the dosage response than the linear model. This showed that most GFI1B-dependent dosage expression responses are better fit by the sigmoid model (median ΔAIC = 18.7, with 70.4% of all trans genes with a significant response having ΔAIC >2, Figure 3D). The responses to dosage modulation of MYB and NFE2 were also better captured by the nonlinearities, but towards less extent (0.14 and 3.4 median ΔAIC, with 20.8% and 40.7% of all trans genes dosage responses having ΔAIC > 2 for MYB and NFE2, respectively, Figure S9A). The broader range of GFI1B expression modulation, providing more data to detect nonlinear trends, likely contributes to this difference. When ignoring those genes classified as unresponsive (genes that their expression did not change upon the TF modulator, see Methods), even more responses of the remaining trans genes were better explained by a sigmoidal model with 83.6%, 26.3% and 63.2% of these having a ΔAIC > 2, for GFI1B, MYB and NFE2 respectively. A similar trend holds even when limiting the models to be fitted to those data points that correspond to a hypothetical one copy loss or gain of the cis gene (Figure S9B), where the median ΔAIC of responsive genes are 7.05, 0.05 and 3.6 for GFI1B, MYB and NFE2 trans responses. Overall this shows that trans responses to TF dosage are dominated by nonlinear behaviours even when the TF dosage changes are not extreme but within biologically plausible ranges.
Gene and transcriptional network properties of dosage response
Utilising a model that effectively captures the variance in our data provided the ability to predict unmeasured TF dosage points and facilitated a direct comparison of trans effects across different cis genes. Employing the sigmoid model (and loess for those with non-monotonic responses), we estimated the continuous expression of trans genes on a uniform fold-change scale across the spectrum of GFI1B, MYB, and NFE2 expression changes (Figure 4A). This estimation was carried out within the empirically observed range of all three cis genes, spanning from log2(FC) −1.83 to 0.51. Subsequent hierarchical clustering of trans gene responses revealed six major clusters of distinct response patterns. For the majority of trans genes, the response to GFI1B and MYB was opposite, with only two small clusters displaying exceptions. Notably, GFI1B generally induced the most substantial response, while NFE2 triggered the smallest range of trans gene response.
Next, we collected diverse annotations for the trans genes to explore the connections between their regulatory properties, disease associations, and selective constraints concerning their response to TF dosage (Figure 4B, C). To quantify these relationships, we assessed significant differences in belonging to these qualitative annotations using the Wilcoxon rank test (Figure 4D) and correlated parameters from the sigmoid model with quantitative gene metrics (Figure 4E). We hypothesised that genes with annotated selective constraint, numerous regulatory elements, and central positions in regulatory networks would exhibit greater robustness to TF changes. Indeed, housekeeping genes demonstrated a considerably smaller dosage response range (Figure 4F). Moreover, genes classified as unresponsive were enriched in the housekeeping category (odds ratio = 2.14, Fisher test p-value = 0.024). The connection between constraint metrics and response properties was also evident in the MYB trans network, where the probability of haploinsufficiency (pHaplo) exhibited a significant negative correlation with the range of transcriptional responses of those trans genes (Figure 4G). However, this result was not reproduced with among the other trans networks (Figure 4E).
While we observed some strong signals on how the response of trans genes similarly vary given similar intrinsic gene properties, most of these differed between GFI1B, MYB and NFE2 trans network responses. We also performed a similar analysis comparing the sigmoid parameters to network properties using the approach in 40, obtaining inconsistent results between TF regulons (Figure S11A, B). This suggests that the link between commonly annotated gene properties and the responses that the genes have are complex and highly context specific, as in our data from a single cell line, they differed between the upstream regulators that were manipulated. Thus, much more data is needed before transcriptional responses can be predicted from gene properties, and conversely to understand the cellular mechanisms that lead to the annotated gene properties.
Nonlinear dosage responses in complex traits and disease
Moving beyond the characterization of mechanisms of transcriptome regulation, a key question is how gradual dosage variation links to downstream cellular phenotypes, and whether these responses exhibit analogous nonlinear patterns. To address this question, we correlated our findings with the expression profiles of various cell types in order to study the myeloid differentiation process, a phenotype well characterised for our K562 model that has been used as a reliable system for investigating erythroid differentiation within myeloid lineages 41 and blood tumours 42. Specifically, leveraging single-cell expression data for bone marrow cell types from the Human Cell Atlas and Human Biomolecular Atlas Project 43, we filtered the expression data to the targeted genes in our study. After aggregating data across donors and normalising expression across cell types (Figure S12A), we compared the expression patterns resulting from each unique transcription factor dosage modulation in relation to each unique cell type expression state. The ensuing correlation can then be construed as a “phenotype,” signifying the similarity between the transcriptional state induced by the TF increase or decrease and the transcriptional state of a specific blood lineage cell type.
Such analyses recapitulate known biology, with GFI1B upregulation 29 and MYB downregulation 44 being crucial factors promoting erythrocyte maturation (Figure 5A). The downregulation of NFE2 instead was negatively related to platelet differentiation. Analysing the correlations as inferred phenotypes suggests potential non-linear relationships (Figure S12B), but these trends should be considered hypotheses that require experimental validation. In summary, this points to cellular phenotypes resulting from gradual TF dosage modulation.
Many of the analysed trans genes are associated with physiological traits and diseases (Figure 4). Understanding the nonlinear trends in the expression of these genes are of particular interest: It helps us comprehend how these genes with physiological impacts may be buffered against upstream regulatory changes, and how their dosage changes as a response to upstream regulators contrasts with genetic variants that contribute to diseases and traits. Additionally, knowing the underlying dosage-to-phenotype curve of a gene can be crucial if this is considered a biomarker for identifying or treating disease. To investigate this, we analysed whether OMIM genes for rare diseases and Mendelian traits or GWAS genes for different blood cell traits (Figure 5B) that are part of the trans networks of genes affected by GFI1B, MYB or NFE2 perturbation are enriched for non-linear dosage responses. As seen in Figure 4, the trans response properties of each gene are highly specific to the regulators and thus analysed in parallel for each cis gene network. An enrichment for nonlinear responses was observed for MYB trans network genes associated with disease and for blood traits related to white blood cells and reticulocytes. These enrichments are particularly interesting given that most trans genes that were sensitive to MYB dosage modulation did not respond with a non-linear trend (Figure S9A).
Despite non-linear responses not being significantly enriched among disease genes across all trans networks, the responses of the same trans gene can show very different dosage responses depending on the upstream regulator being tuned. In Figure 5C we show a few examples of genes associated with diseases (1 or more disease phenotypes 45): FOXP1 is a haploinsufficient and potentially triplosensitive transcription factor associated with intellectual disability; yet it shows a strong response especially to GFI1B dosage across a wide range. NF1A is a haploinsufficient developmental disorder gene with a similar response pattern.
However, it is difficult to interpret their expression response in K562 cells when their most apparent phenotypic effects likely derive from other cell types. RHB is the Rhesus blood type gene where a common deletion of the gene causes the Rh-blood type in homozygous individuals, with a strong nonlinear response to GFI1B levels. A particularly interesting gene is TUBB1, part of β-tubulin, that causes autosomal dominant macrothrombocytopenia or abnormally large platelets. Here, K562 cells are a reasonable model system, being closely related precursors to megakaryocytes that produce platelets. Interestingly, GFI1B loss also causes a macrothrombocytopenia phenotype in mice 46, and in our data TUBB1 expression decreases quickly as a function of decreased GFI1B expression but then plateaus at a level that corresponding to loss of one copy of TUBB1. This raises the hypothesis that low GFI1B levels may cause macrothrombocytopenia at least partially via reducing TUBB1 expression.
Discussion
In this paper, we have investigated how gradual dosage modulation of transcription factors contributes to dosage-sensitive transcriptional regulation and investigated its potential phenotypic consequences. First, we set up an easily scalable and generalizable CRISPRi/CRISPRa approach with tiling sgRNAs for gradual titration of gene expression, with reagents that can be designed with data only of the TSS and easily ordered at scale. Alternative approaches that rely e.g. on targeting CREs that are often unknown, dramatic overexpression, or laborious setting up of constructs for each gene are less practical for large-scale analyses. Our approach appears best suited for expression modulation in the biologically reasonable range, and other methods would be needed for dramatic overexpression or complete silencing of the target genes. Our inability to substantially increase MYB expression indicates the need for further work and larger cis gene sets to fully understand how widespread this is and to which extent this depends on cis-regulatory properties versus feedback and buffering mechanisms. Nevertheless, we believe that the approach proposed here is a useful complement to the diversifying set of tools for dosage modulation for different purposes 18–22.
In this work, we made use of targeted transcriptome sequencing to avoid complications from the sparsity of single cell data. While highly accurate targeted readout of the cis gene expression linked to each sgRNA is a core component of our approach, analysing trans responses could also be achieved by standard single cell sequencing of the full transcriptome, possible in combination with a targeted readout of transcripts of particular interest. In this study, the targeted genes were selected based on prior data of responding to GFI1B, NFE2 or MYB regulation and thus do not represent an unbiased or random sample of genes. An interesting future extension would be the addition of single-cell protein quantification to confirm that the detected mRNA levels correspond to protein levels, but this remains technically challenging.
Our results show that nonlinear responses to gradual up- and downregulation of TF dosage are widespread and can be detected even without extreme overexpression or full knockout of the TF. The patterns of transcriptional responses are highly context-specific and vary between upstream regulators. Further work with larger sets of cis and trans genes as well as direct quantification of cellular readouts will be needed to fully characterise the patterns and mechanisms of downstream impacts on gene dosage. However, our findings indicate important directions for future research. Firstly, the widespread nonlinearity suggests that interpolation or extrapolating gene function assessments from classical molecular biology approaches with drastic knock-outs or knock-downs may have limitations, as their effects can be quantitatively and qualitatively different than more modest perturbations that typically occur in nature. This may be particularly relevant for essential and highly dosage-sensitive genes, where applying our gradual dosage modulation framework can provide opportunities for functional characterization at perturbation levels that do not kill the cells. Secondly, we show that the effects of up- and downregulation are qualitatively and quantitatively different, which calls for increased attention to analysing both directions of effect, which also occur in natural responses to genetic variants and environmental stimuli.
Gene dosage sensitivity has typically been studied by human genetics and genomics methods 47–49. The experimental approach pursued in this study and the computational approaches are fundamentally different and complement each other: while human genetics is powerful for capturing functional importance on physiological phenotypes via patterns of contemporary population variation and selective constraint, experimental approaches provide more granularity and insights into cellular mechanisms. Furthermore, while the convergence of disease effects of common and rare variants affecting the same gene is a well-known phenomenon 6,7, the sparsity of variants makes it difficult to properly model allelic series as a continuous dosage-to-phenotype function for individual genes. Experimental approaches can provide a powerful complement to this. Altogether, we envision that combining these perspectives into true systems genetics approaches will be a powerful way to understand how gene dosage variation contributes to human phenotypes from molecular to cellular and eventually physiological levels.
Acknowledgements
This work was funded by NIH grants R01MH106842, R01AG057422, DP2HG010099, R01HG012790, and R01GM122924; a grant from the Knut and Alice Wallenberg Foundation to SciLifeLab for research in Data-driven Life Science, DDLS (KAW 2020.0239); funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (Grant agreement No. 101043238); a European Molecular Biology Organization Postdoctoral Fellowship (ALTF 345-2021) to J.D.; a Canadian Institutes of Health Research Banting Postdoctoral Fellowship and NIH/NHGRI (K99HG012792) to J.A.M.
Additional information
Competing Interests
J.D. is CEO and co-founder with equity in Allostery Exploration Technologies, S.L. T.L. was a paid advisor to GSK, and is an advisor and has equity in Variant Bio.
Author Contributions
J.D. and T.L. conceived the study. J.D. performed the experiments, with contributions from
J.A.M. and M.Z.. N.S. contributed experimental resources to the study. J.D. performed the computational analyses, with contributions from M.M., S.G. and J.A.M.. J.D. and T.L. wrote the manuscript with contributions and review from all the authors.
Data availability and materials
All code used in this study is available at https://github.com/LappalainenLab/d2n_ms. Raw sequencing data has been submitted to GEO (accession number GSE257547).
Experimental methods
Plasmids, bacteria strains and cell lines used
Plasmids
pCC_05: Lentiviral Puromycin CRISPRa dCas9-VPR system (Addgene 139090)
pGC02: Lentiviral Blasticydin plasmid with CRISPRi KRAB-dCas9-MeCP2 system (Addgene 170068)
pJDE003: Lentiviral Blasticydin CRISPRa dCas9-VPR system (this study)
pGC03: Lentiviral Puromycin sgRNA library cloning vector (Addgene 170069)
pMD2G: Lentiviral envelope plasmid (Addgene 12259)
psPAX2: Lentiviral packaging plasmid (Addgene 12260)
Bacteria E.coli strains
5-alpha competent cells (NEB C2987H)
One Shot Stbl3 competent cells (Invitrogen 1934665)
Endura electrocompetent cells (Lucigen 60242-2)
Cell lines:
HEK293FT (Thermo Fisher Scientific R70007); cells were maintained at 37℃ and 5% CO2 in high glucose DMEM (Cytiva SH30022.01) supplemented with 10% Serum Plus II (Sigma-Aldrich 14009C)
K562 (ATCC, CCL243);
Cells were maintained at 37℃ and 5% CO2; HEK293FT were cultured in high glucose DMEM (Cytiva SH30022.01) supplemented with 10% Serum Plus II (Sigma-Aldrich 14009C); K562 were cultured in IMDM, GlutaMAX (Thermo Scientific 31980097) supplemented with 10% Serum Plus II.
CRISPRa vector construction
To construct the vector harbouring the CRISPRa system (pJDE003), the CRISPRi (KRAB-dCas9-MeCP2) gene fusion of pGC02 1 was replaced with dCas9-VPR cassette, which was PCR amplified (Q5 High-Fidelity 2X Master Mix, NEB M0492L) from the plasmid pCC_05 2 with primers oJDE005 and oJDE006 following instructions from manufacturer. pGC02 was digested with XbaI-FD and BamHI-FD (Thermo Fisher FD0685 and FD0054) and sequentially dephosphorylated with FastAP (Thermo Fisher EF0651) following the manufacturer’s recommendations. The digested pGC02 vector and the PCR insert with the CRISPRa system (previously treated with a 15 min DpnI enzyme incubation, Thermo Fisher FD1704) were assembled by Gibson assembly using a 2:1 insert:vector ratio with Gibson Assembly Master Mix (NEB E2611S). Assemblies were transformed into NEB 5-alpha E.coli competent cells and single colonies were picked and sequence validated by Sanger sequencing. Frozen stock of the correct construct cells were regrown for plasmid Maxiprep extraction (QIAGEN 12362) for subsequent virus production.
CRISPRa K562 cell line construction and functional validation
Lentivirus was produced by polyethylenimine linear MW 25000 (Polysciences 23966) transfection of HEK293FT cells with the transfer plasmid containing a Cas9-VPR effector, packaging plasmid psPAX2 (Addgene 12260) and envelope plasmid pMD2.G (Addgene 12259). After 72 h post-transfection, cell media containing lentiviral particles was harvested and filtered through 0.45 μm filter Steriflip-HV (Millipore SE1M003M00). One volume of Lentivirus Precipitation Solution (Alstem VC100) was added to the collected lentivirus, mixed and stored overnight at 4C. The mix was centrifuged for 30 min at 1,500g, and the pellet of lentiviral particles were resuspended in 1/10th of the original volume of DMEM media. Lentivirus vials were frozen at −80C and later thawed for transduction.
To construct the monoclonal K562 cell line with the CRISPRa system, the dCas9-VPR lentivirus was transduced into one million K562 cells using 100 µl of 10X concentrated lentivirus in a total volume of 1 ml (high MOI). After 24 hours, the media was replaced with fresh IMDM, and 48 hours after transduction, blasticidin (A.G. Scientific B-1247) was added to a final concentration of 10 µg/µl for 16 days. Monoclonal cell lines were sorted by FACS (Sony Cell Sorter SH800) into a 96-well plate. The presence of dCAS9 protein in several growing clones was confirmed by western blot (Primary antibody: Purified anti-CRISPR CAS9 antibody; BioLegend 844302. Secondary antibody: LI-COR 925-32212) and protein levels were normalised to GAPDH (Primary antibody: GAPDH (14C10) Rabbit; Cell Signalling Technology 2118S. Secondary antibody: LI-COR 925-68073).
To select the final monoclonal CRISPa cell line, the three clones with the highest protein expression in the western blot were subjected to functional validation to test for activation activity. Lentiviral guides designed from 2 targeting CD4 (Anti-CD4 Mouse Monoclonal Antibody (FITC), BioLegend, 300505), which is lowly expressed in K562, CD19 (Anti-CD19 Mouse Monoclonal Antibody (APC),BioLegend, 302211) with null expression, and CD45 (PE anti-human CD45, BioLegend, 368509) with intermediate expression, were independently transduced into all three monoclonals, and after puromycin selection, the expression of this markers was screened by FACS at day 4 and at day 10 or 11 after transduction. The clone with the strongest and most consistent activity was selected.
Gene selection for targeted sequencing and design of probe custom panel
The four selected dosage genes were GFI1B, NFE2, MYB, and TET2. GFI1B and NFE2 were chosen due to their reported trans effects following the inhibition of their cis-CREs 3. MYB was selected for being downstream of the GFI1B network, and TET2 was selected as an unrelated gene to those transcriptional networks. Both MYB and TET2 qualified as oncogene and tumour suppressor functions in K562 (https://depmap.org/portal/), making them ideal choices to determine the impact of growth effects on the experiment.
The additional 88 genes captured by targeted sequencing were selected based on the significant trans effects of GFI1B and NFE2 inhibition from 3 including two control genes, GAPDH, and LHX3 genes, one highly constantly expressed “housekeeping gene” and the other with no reported expression in K562 cell lines, respectively. The remaining 86 genes were selected based on Morris et al. to include 1) 29 genes that overlapped between the NFE2 and GFI1B network, 2) 47 trans genes for GFI1B, 10) trans genes for NFE2. The number of trans genes selected from each unique network was proportionally chosen, given the size of each trans network, and oversampling TFs and TF targets as defined in 4, as well as maintaining the proportional co-expression cluster structure as defined in 3 (Figure S1a,b). Additional filters in the selection included a minimum expression of 0.1 mean UMI/cell and the lack of alternative 5’ splice isoforms and a unique Ensembl ID.
The 10X Probe Full Custom Panel design tool was used to design the targeted gene expression probe library. A total of 687 probes (∼15%) were discarded because they covered transcript regions with a median coverage per base of < 3 reads/bp (for medium and highly expressed genes) or <1 reads/bp (for lowly expressed genes). All probes for LHX3 (0 median reads/bp) were retained. In addition, 93 probes covering the entire transcript sequence of the dCas9-VPR and KRAB-dCas9-MeCP2 transcript were included, resulting in a final total of 4,405 probes. The xGen™ Custom Hybridization Capture Panel of biotinylated oligos was ordered and synthesised at IDT.
Gene dosage sgRNA library design and cloning
The sgRNA library contained a total of 96 guides (51 tiling, 8 TSS, 20 attenuated, 12 enhancer and 5 non-targeting controls). All guides were designed to not contain the U6 terminator sequence, repeats of five or more consecutive G, C or As, as well as not falling in the genomic region where K562 cell line has alternative alleles compared to the human genome reference (Hg38). All guides were scored with FlashFry 5 to obtain off-target and on-target activity scores that allowed the selection of the best scoring guides. Tiling guides were designed to target different regions of the promoter, TSS and beginning of the gene body of each dosage gene, spanning a total average distance of 1400 bp (TSS in the centre), each being on average distant from one another of 110 bp. The sequences of the two TSS guides were obtained from 6. The sgRNAs targeting enhancers were picked from previously reported work that showed a CRISPR-based evidence of enhancer activity (GFI1B 1, NFE2 7, MYB 8). The five attenuated guides for each gene were manually designed following the rules described in 9 to span a range of activities, including a single point mutation on the best scoring guide that targeted the TSS.
Overhangs with homology regions to the pGC03 plasmid (18bp downstream and 22 bp upstream) were added to the sgRNA sequence to be able to directly clone the ssDNA oligos into the plasmid. The 96 sgRNAs were ordered in IDT as single stranded DNA oligos (total 60bp) in a 96 well-plate to 100 pmol scale. The oligos were pooled at equimolar concentration and diluted to a final concentration of 0.2 uM. The library was cloned into the BsmBI digested plasmid pGC03 using 10 reactions of the NEBuilder HiFi DNA Assembly kit following the manufacturer’s instructions. All the reactions were pooled and the DNA precipitated using Isopropanol, GlycoBlue (Thermo Scientific AM9515) and 50mM NaCl for 15 min at RT. Following two washes with Ethanol 70%, the assembly was resuspended with 15ul of 0.2X TE.
To transform the library into E.coli, 1ul of the assembly was mixed with 25 uL of Endura cells under manufacturer’s electroporation conditions, then plated onto 245 x 245 mm square LB 100 ug/ml Carbenicillin plates. The plates were grown ON, and >2.5e5 transformants were obtained, ensuring the complexity of the library was maintained at >1000 cells per unique sgRNA. All colonies were collected and subjected to maxiprep using the Maxi Fast-Ion Plasmid Kit, Endotoxin Free kit (IBI Scientific IB47123). The representation of the library was assessed through MiSeq shallow sequencing (Illumina).
sgRNA lentiviral library production and cell culture assay
The lentiviral library was produced by transfecting ∼80 million HEK293FT cells with a transfer plasmid containing the 96 sgRNA library, along with the packaging plasmid psPAX2 and envelope plasmid pMD2.G, using polyethylenimine linear MW 25000. The supernatant media was replaced with fresh D10 10% BSA six hours after transfection, and the virus was collected and filtered through 0.45 µm filters after 48 hours. The lentiviral library was then concentrated 2X using the Precipitation Lentiviral Solution, aliquoted, and stored at −80°C for subsequent transduction.
Both CRISPRi and CRISPRa K562 cell lines were independently transduced with different titers of the lentiviral sgRNA library at a low MOI (one sgRNA per cell). Twenty hours post-transduction, the cell media was replaced with fresh IMDM 10% Serum Plus II Blasticidin 5 µg/mL, and four hours later, Puromycin (Invivogen ant-pr-1) was added at a final concentration of 2 µg/mL to select for cells with sgRNA integration. The transduction batch with an infection rate of ∼10% was selected, and cells were sorted to near purity using FACS to remove dead cells. Cells were maintained at >90% survival and a maximum confluency of 700,000 cells/mL. On day 8 post-transduction, the cells were collected and prepared for cell hashing.
Multimodal single-cell experiment and targeted sequencing
Cell hashing was performed as previously described using four hashtag-derived oligonucleotides (HTOs) in a hyperconjugation protocol 10. Each transduced cell line was split into four batches of 500,000 cells, resulting in a total of 8 different hashes. After incubation and washes, all 8 hashed batches were pooled together and run in two reaction lanes of the 10X Chromium Next GEM Single Cell 5’ Reagent Kit v2 (single indexing, PN-1000265 and PN-1000190). The manufacturer’s protocol was followed with modifications stipulated in the ECCITE-seq protocol 11. For each GEM reaction, 42,000 cells from the hash pool were used to obtain approximately 21,000 total cells, including “multiplets” (multiple cells per droplet counts). Gene expression (cDNA), hashtags (HTOs), and guide RNA (Guide-derived oligos, GDOs) libraries were constructed following the 10x Genomics and ECCITE-seq protocols (https://cite-seq.com/eccite-seq/) with minimal modifications.
Specifically, the antibody pool protein tag library steps were ignored, and a custom-designed probe library was used to enrich the cDNA for the genes of interest in the 10X Targeted Gene Expression protocol (10X PN-1000248). The resulting libraries were sequenced using an Illumina Nextseq 500/550 Mid-Output v2.5 Kit (150 cycles). The targeted enrichment of the dCas9 transcripts was performed separately using an independent probe library and was sequenced together with additional HTO libraries using the Illumina Miseq Reagent Kit v3 (150 cycles).
Computational and statistical analyses
From fastqs to QCed and demultiplexed UMI normalised matrices
FastQC was used to demultiplex the different samples of the three different modalities from the different 10X chip lanes, which each was processed independently. For the cDNA modality, the UMI count matrix was obtained using Cellranger count, including the targeted-panel argument to get the additional filtered matrices and summary statistics. Cells with less than 500 UMIs per cell or less than 50 genes with at least 1 UMI per cell were discarded. The top 1% cells containing the highest UMI content were also discarded. The expression of all genes across 10X lanes were extremely reproducible (Pearson r = 0.999), showing a ∼5-fold UMI count increase in contrast to the non-targeted transcriptome (Figure S1c).
For the GDO modality (sgRNAs), Cellranger count was also used using the CRISPR Guide Capture Analysis mode, which uses a Gaussian mixture model to call sgRNA per cell. The cells containing more than one sgRNA were discarded.
To classify each cell into their corresponding CRISPR system of origin (CRISPRi or CRISPRa), both the HTO modality (protein hashes) and the expression of the dCas9 targeted transcript was used. Protein hashes were called using Alevin salmon 12, and the resulting HTO UMI matrix was mixed in with the cDNA matrix containing the expression of the CRISPRi and CRISPRa genes. This matrix was normalised and scaled using Seurat v4 13 and used to generate a UMAP based on the expression of protein hashes and the dCas9 transcripts expression. Clusters were identified and manually assigned to an HTO category given the expression pattern of each cluster. Finally, the 5% cells classified as CRISPRi that had the lowest expression of CRISPRi transcript were discarded, as well as those 5% of CRISPRa classified cells that had the highest CRISPRi transcript expression. In total, 20,001 (10,647 CRISPRi and 9,354 CRISPRa) cells passed all filters and were used for subsequent analyses.
Once each single cell was classified into a unique sgRNA perturbation and to a cell line of origin, the cDNA UMI matrices of the two 10X lanes were merged and afterwards normalised using a the log1p normalisation method of Seurat’s NormalizeData (Seurat version 4.3). On average, each unique sgRNA perturbation was measured in 81 and 86 cells for the CRISPRa and CRISPRi, respectively (Figure S1d)
Expression fold-change calculation and non-target sgRNA filtering
As estimates of changes in expression, we used a pseudo-bulk differential analyses approach. To get rid of the batch effects deriving from each cell line (CRISPRa vs. CRISPRa) (Figure S3a), for each unique perturbation we calculated log2 fold-change of the of the expression a gene against the expression of that gene in the population of cells harbouring the NTC sgRNAs of that particular cell line. We used Seurat’s FindMarkers function to calculate the log2FC and the p-values of a Wilcoxon Rank Sum test.
Before running the differential analyses on all targeted genes, across all unique CRISPR perturbations we identified those NTC sgRNAs that had potential unexpected off target activity and thus could not be used as negative controls. For all possible unique NTC sgRNA pairs we run the above differential expression analysis on all 92 targeted genes. We discarded NTC sgRNAs that showed more than one DE gene (FDR 0.05) in more than one pairwise comparison, and the differential genes showed consistent patterns of change in expression. For this reason, cells harbouring sgRNA NTC_2 on the CRISPRa modality were discarded, as this particular perturbation showed consistent undesired activation of PPP1R14A and CTCFL genes. Additionally, we ran Sceptre 14 using the resulting group of control cells to validate that our samples were calibrated correctly (Figure S1e).
Once those potential outlier NTCs were discarded, the log2FC of each targeted gene in each unique sgRNA and cell line condition was calculated. Adjusted FDR p-values were calculated across all tests to later on call significance on DE genes. The obtained fold-changes and FDRs were used for all subsequent analyses.
Linear, loess and sigmoidal model fitting
To identify the best predictive model of each cis-gene dosage to trans-gene fold-change, we fitted three types of models to the data: linear (using the R lm function), a four parameter sigmoid (using the drm(fct = L.4()) function from the R dcr package) and a LOESS fit (R loess function). To evaluate and compare the goodness of fit of the linear vs. the sigmoid model taking into account overfitting, we calculated the Akaike information criterion (AIC) using the AIC function from the R stats package.
To obtain an accurate prediction of each trans gene expression given TF dosage and avoid overfitting, a 10-fold cross validation scheme was followed by fitting the sigmoid model individually to each curve. The data was randomly split in 10 groups, where 90% of the data was used for training and the remaining 10% for testing. To obtain the values of each individual sigmoid fit for each dosage and trans gene response, the average and standard deviation of each parameter value was calculated across the 10 trained models.
Those trans genes with a slope significantly different from 0 (FDR adjusted p-value of a z-test across the 10 fold-CV parameter outputs) and with a min-to-max range significantly higher than 0.05 (FDR adjusted p-value of a z-test across the difference between the min and max asymptotes parameters in the 10 fold-CV), were classified as as “responsive” genes. The remaining genes were classified as “unresponsive”. The top 5% trans genes of the GFI1B trans network with the largest ΔRMSE between the LOESS fit and the sigmoid fit (RMSESigmoid - RMSELOESS) were classified as non-monotonic and the curve trend manually validated. For the five trans genes classified to have a non-monotonic gene expression response, their predicted expression upon TF dosage change was calculated using the LOESS model instead of the sigmoid one.
Gene-specific properties
Diverse gene annotations and properties were collected to compare with the different trans genes response properties (related to Figure 4). Quantitative annotations included the gene biotype(Ensembl 15), Housekeeping genes 16, transcription factors 4, genes associated with at least one disease (OMIM 17) and genes associated with blood-related complex traits (obtained from 3).
Quantitative features included the probability of being loss-of-function intolerant scores (pLI)18 and synonymous and missense Z scores (mis z) 18,19, which were obtained from the GnomAD database. Haploinsufficiency probability scores were obtained from 20. To obtain the number of ChIP-Seq peaks of a cis gene within the promoter region of trans-genes (n peak [cis gene]), we utilised the regulon generated by Minaeva et al. 2024 21. This regulon was created by mapping transcription factor peaks to transcription start sites (TSS) of the 50% expressed isoforms for each gene in K562 cells, with subsequent application of a ±1 Kb proximity filter. Mean expression of genes from bone marrow cell types were obtained from Hay et al. 2018 22 and averaged across donors. The number of protein-protein interactions of each gene within the entire human proteome (Num PPIs1) was obtained from the STING database 23.
To test significant differences between groups of genes (qualitative features), the Wilcoxon rank test was used. For quantitative features, Pearson correlation between parameters from the sigmoid model with quantitative gene metrics was used. Non-responsive and non-monotonic genes in each trans network were excluded.
Code and data accessibility
All code used in this study is available at https://github.com/LappalainenLab/d2n_ms. Raw sequencing data has been submitted to GEO (accession number GSE257547).
Note
This reviewed preprint has been updated to use the correct text; previously, a version prior to the one reviewed was presented here.
References
- 1.Systematic localization of common disease-associated variation in regulatory DNAScience 337:1190–1195
- 2.The GTEx Consortium atlas of genetic regulatory effects across human tissuesScience 369:1318–1330
- 3.A brief history of human disease geneticsNature 577:179–189
- 4.Discovery of target genes and pathways at GWAS loci by pooled single-cell CRISPR screensScience 380
- 5.Mendelian inheritance revisited: dominance and recessiveness in medical geneticsNat. Rev. Genet 24:442–463
- 6.Exome sequencing and analysis of 454,787 UK Biobank participantsNature 599:628–634
- 7.Phenotype-Specific Enrichment of Mendelian Disorder Genes near GWAS Regions across 62 Complex TraitsAm. J. Hum. Genet 103:535–552
- 8.Modified penetrance of coding variants by cis-regulatory variation contributes to disease riskNat. Genet 50:1327–1334
- 9.Polygenic background modifies penetrance of monogenic variants for tier 1 genomic conditionsNat. Commun 11
- 10.An allelic-series rare-variant association test for candidate-gene discoveryAm. J. Hum. Genet 110:1330–1342
- 11.Genome-scale CRISPR pooled screensAnal. Biochem 532:95–99
- 12.A cross-disorder dosage sensitivity map of the human genomemedRxiv
- 13.Gene expression analysis identifies global gene dosage sensitivity in cancerNat. Genet 47:115–125
- 14.Identification and characterization of essential genes in the human genomeScience 350:1096–1101
- 15.High-Resolution CRISPR Screens Reveal Fitness Genes and Genotype-Specific Cancer LiabilitiesCell 163:1515–1526
- 16.Parallel genome-scale loss of function screens in 216 cancer cell lines for the identification of context-specific genetic dependenciesSci Data 1
- 17.Massively Parallel Interrogation of the Effects of Gene Expression Levels on FitnessCell 166:1282–1294https://doi.org/10.1016/j.cell.2016.07.024
- 18.Titrating gene expression using libraries of systematically attenuated CRISPR guide RNAsNat. Biotechnol 38:355–364
- 19.Mismatch-CRISPRi Reveals the Co-varying Expression-Fitness Relationships of Essential Genes in Escherichia coli and Bacillus subtilisCell Syst 11:523–535
- 20.CasTuner is a degron and CRISPR/Cas-based toolkit for analog tuning of endogenous gene expressionNat. Commun 14
- 21.Dose-dependent activation of gene expression is achieved using CRISPR and small molecules that recruit endogenous chromatin machineryNat. Biotechnol 38:50–55
- 22.Dissecting reprogramming heterogeneity at single-cell resolution using scTF-seqbioRxiv https://doi.org/10.1101/2024.01.30.577921
- 23.Spurious regulatory connections dictate the expression-fitness landscape of translation factorsMol. Syst. Biol 17
- 24.Precise modulation of transcription factor levels identifies features underlying dosage sensitivityNat. Genet https://doi.org/10.1038/s41588-023-01366-2
- 25.Dosage amplification dictates oncogenic regulation by the NKX2-1 lineage factor in lung adenocarcinomabioRxiv https://doi.org/10.1101/2023.10.26.563996
- 26.Systematic differences in discovery of genetic effects on gene expression and complex traitsNat. Genet 55:1866–1875
- 27.Deregulated Regulators: Disease-Causing cis Variants in Transcription Factor GenesTrends Genet 36:523–539
- 28.Systematic Functional Dissection of Common Genetic Variation Affecting Red Blood Cell TraitsCell 165:1530–1545
- 29.From cytopenia to leukemia: the role of Gfi1 and Gfi1b in blood formationBlood 126:2561–2569
- 30.Whole-Exome Sequencing Identifies Loci Associated with Blood Cell Traits and Reveals a Role for Alternative GFI1B Splice Variants in Human HematopoiesisAm. J. Hum. Genet 99:481–488
- 31.Altered NFE2 activity predisposes to leukemic transformation and myelosarcoma with AML-specific aberrationsBlood 133:1766–1777
- 32.B-myb is an essential regulator of hematopoietic stem cell and myeloid progenitor cell developmentProc. Natl. Acad. Sci. U. S. A 111:3122–3127
- 33.Rational design of highly active sgRNAs for CRISPR-Cas9-mediated gene inactivationNat. Biotechnol 32:1262–1267
- 34.Optimized sgRNA design to maximize activity and minimize off-target effects of CRISPR-Cas9Nat. Biotechnol 34:184–191
- 35.FlashFry: a fast and flexible tool for large-scale CRISPR target designBMC Biol 16
- 36.Optimized libraries for CRISPR-Cas9 genetic screens with multiple modalitiesNat. Commun 9
- 37.High-Throughput Screens of PAM-Flexible Cas9 Variants for Gene Knockout and Transcriptional ModulationCell Rep 30:2859–2868
- 38.New developments on the Encyclopedia of DNA Elements (ENCODE) data portalNucleic Acids Res 48:D882–D889
- 39.KRAB-Zinc Finger Proteins: A Repressor Family Displaying Multiple Biological FunctionsCurr. Genomics 14:268–278
- 40.Specifying cellular context of transcription factor regulons for exploring context-specific gene regulation programsbioRxiv https://doi.org/10.1101/2023.12.31.573765
- 41.Dynamic transcriptomes of human myeloid leukemia cellsGenomics 102:250–256
- 42.Matching cell lines with cancer type and subtype of origin via mutational, epigenomic, and transcriptomic patternsSci Adv 6
- 43.Single-cell multiomic analysis identifies regulatory programs in mixed-phenotype acute leukemiaNat. Biotechnol 37:1458–1465
- 44.Down-regulation of Myc is essential for terminal erythroid maturationJ. Biol. Chem 285:40252–40265
- 45.org: Online Mendelian Inheritance in Man (OMIM®), an online catalog of human genes and genetic disordersNucleic Acids Res 43:D789–98
- 46.Gfi1b controls integrin signaling-dependent cytoskeleton dynamics and organization in megakaryocytesHaematologica 102:484–497
- 47.A structural variation reference for medical and population geneticsNature 581:444–451
- 48.Genetic regulatory variation in populations informs transcriptome analysis in rare diseaseScience 366:351–356
- 49.An RNA-informed dosage sensitivity map reflects the intrinsic functional nature of genesAm. J. Hum. Genet 110:1509–1521
- 1.Discovery of target genes and pathways of blood trait loci using pooled CRISPR screens and single cell RNA sequencingbioRxiv https://doi.org/10.1101/2021.04.07.438882
- 2.High-Throughput Screens of PAM-Flexible Cas9 Variants for Gene Knockout and Transcriptional ModulationCell Rep 30:2859–2868
- 3.Discovery of target genes and pathways at GWAS loci by pooled single-cell CRISPR screensScience 380
- 4.Benchmark and integration of resources for the estimation of human transcription factor activitiesGenome Res 29:1363–1375
- 5.FlashFry: a fast and flexible tool for large-scale CRISPR target designBMC Biol 16
- 6.Mapping information-rich genotype-phenotype landscapes with genome-scale Perturb-seqCell 185:2559–2575
- 7.Global Analysis of Enhancer Targets Reveals Convergent Enhancer-Driven Regulatory ModulesCell Rep 29:2570–2578
- 8.Regulation of MYB by distal enhancer elements in human myeloid leukemiaCell Death Dis 12
- 9.Titrating gene expression using libraries of systematically attenuated CRISPR guide RNAsNat. Biotechnol 38:355–364
- 10.Simultaneous epitope and transcriptome measurement in single cellsNat. Methods 14:865–868
- 11.Multiplexed detection of proteins, transcriptomes, clonotypes and CRISPR perturbations in single cellsNat. Methods 16:409–412
- 12.A Bayesian framework for inter-cellular information sharing improves dscRNA-seq quantificationBioinformatics 36:i292–i299
- 13.Integrated analysis of multimodal single-cell dataCell 184:3573–3587
- 14.SCEPTRE improves calibration and sensitivity in single-cell CRISPR screen analysisGenome Biol 22
- 15.Ensembl 2023Nucleic Acids Res 51:D933–D941
- 16.Human housekeeping genes, revisitedTrends Genet 29:569–574
- 17.org: Online Mendelian Inheritance in Man (OMIM®), an online catalog of human genes and genetic disordersNucleic Acids Res 43:D789–98
- 18.Analysis of protein-coding genetic variation in 60,706 humansNature 536:285–291
- 19.A framework for the interpretation of de novo mutation in human diseaseNat. Genet 46:944–950
- 20.A cross-disorder dosage sensitivity map of the human genomeCell 185:3041–3055
- 21.Specifying cellular context of transcription factor regulons for exploring context-specific gene regulation programsbioRxiv https://doi.org/10.1101/2023.12.31.573765
- 22.The Human Cell Atlas bone marrow single-cell interactive web portalExp. Hematol 68:51–61
- 23.STRING v10: protein-protein interaction networks, integrated over the tree of lifeNucleic Acids Res 43:D447–52
Article and author information
Author information
Version history
- Preprint posted:
- Sent for peer review:
- Reviewed Preprint version 1:
Copyright
© 2024, Domingo et al.
This article is distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use and redistribution provided that the original author and source are credited.
Metrics
- views
- 112
- downloads
- 0
- citations
- 0
Views, downloads and citations are aggregated across all versions of this paper published by eLife.