Abstract
Gene regulatory networks are now at the forefront of precision biology, which can help researchers better understand how genes and regulatory elements interact to control cellular gene expression, offering a more promising molecular mechanism in biological research. Interactions between the genes and regulatory elements involve different promoters, enhancers, transcription factors, silencers, insulators, and long-range regulatory elements, which occur at a ∼10 µm nucleus in a spatiotemporal manner. In this way, three-dimensional chromatin conformation and structural biology are critical for interpreting the biological effects and the gene regulatory networks. In the review, we have briefly summarized the latest processes in three-dimensional chromatin conformation, microscopic imaging, and bioinformatics, and we have presented the outlook and future directions for these three aspects.
Keywords: Gene regulatory network (GRN), Chromatin conformation, Microscope imaging, Regulatory elements
1. Introduction
A basic problem in biology is how a fertilized egg develops into a creature that consists of hundreds of different cell types. Decades of research have revealed that oosperm development is a complex process that involves spatiotemporal specific gene expression as well as regulatory element modulation [1], [2]. Researchers have discovered that the genome structure is not one-dimensional from the initial central dogma to this post-genomics [3], [4]. Instead, it more likely has a complicated three-dimensional organization in the nucleus [5]. Concurrently, with the tremendous progress made in epigenetics, molecular mechanisms, and non-coding sequences of transcriptional regulation [6], [7], [8], [9], [10], massive research have found that biological phenotypes are regulated by the gene regulatory network (GRN), which is encompassed by multi-genes, regulatory elements, transcription factors (TFs) and co-factors. A gene is not only modulated by one gene, TF, or regulatory element, it tends to be regulated by multiple components of them. Moreover, recent studies found that the chromatin architecture could profoundly affect the GRN [11] (Fig. 1). For example, the folding of chromatin provides a distal regulatory element the chance to interact with its target genes and then regulate the expression of genes. Thus, proper analysis of the GRN is pivotal in studying the correlation between gene expression and biological phenotypes [12], [13], [14].
Fig. 1.
Diagram of the Gene Regulatory Network. (A) schematically, the gene regulatory network. (B) the diagram for the 3D architecture of the gene regulatory network. Note: Gene A is regulated by the gene B and regulatory element; the gene B is regulated by gene A, gene C, and TF2; the gene C is regulated by the gene B, gene A, and TF2; and the gene D is regulated by the gene B and TF1.
The focus of gene regulatory networks is to investigate the transcription factors and signaling components involved in developmental processes in which the chromatin architecture in the nucleus is established by transcription participators in essence. At present, some progress has been made in analyzing GRNs using a variety of methods. Transcriptome analysis is a prevalent technology for GRN analysis, which studies gene expression changes in a dynamic state. The network is inferred via a bioinformatics mode. There is room to improve its reliability and reproducibility. For instance, the noise in the transcriptome has a profound influence on classification accuracy. With the most recent advances in three-dimensional (3D) genomics and the relationship between GRNs and 3D structure, 3D and structural biological data have been integrated with the transcriptome and other -omics methods, greatly improving the interpretation of GRNs and their biological effects. In this review, we summarized the recent progress of these aspects as well as the advantages and limitations of the techniques used in analyzing the GRNs. We also illustrated some points for the future analysis of GRNs.
2. Construction of a gene regulatory network supported by sequence-based technology
Studying the interaction between chromatin, regulatory elements, and transcription factors is an important part of the transcriptional regulatory network. Characterizing various components of transcriptional regulation, such as promoters, enhancers, and super-enhancers (SEs) and their 3D interactions is a standard method for interpreting transcriptional regulation [15], [16]. The specific expression of genes is hightly determined by regulatory elements, and enhancers are the important one of regulatory elements [17]. Genome-wide enhancers and SEs can be identified by sequence-based techniques (such as STARR-seq) [6], ChIP-seq (such as H3K27ac, H3K4me1, p300) [18], [19] and chromatin accessibility (ATAC-seq, DNase-seq) [20], [21]. Among them, SEs are specific regulators of transcriptional regulatory networks that heavily interact with multiple genes and other elements involved in cell identity. According to the report, SEs are considered large clusters of regulatory elements and possess stronger abilities to function in GRNs compared with typical enhancers [22]. Research in humans found that the super-enhancers in the retina are enriched for two TFs (CRX and NRL) involved in photoreceptors, and genes that overlap with SEs are retinal-specific and have a higher expression [23]. The binding of gene and tissue-specific transcription factors is a key component and driver of gene expression regulation [24]. Another important issue with complex regulatory networks is the superposition effect of multiple enhancers. In regulatory networks, each enhancer is independent and functions in a synergistic manner [25], [26]. The study of the Indian hedgehog (Ihh) gene showed that the Ihh gene was regulated by nine specific enhancers in a synergistic manner, and that gene expression was up-regulated with the increase in enhancer copy number [27]. In the study of the PIM1 gene in human cells, it was found that the inhibition of this gene must simultaneously knock out several enhancers rather than a single enhancer [28]. Therefore, an important factor in gene transcription regulation is composed of complex enhancer sets, which together achieve fine regulation of genes. Nevertheless, the distance between the enhancer and the promoter is not a major factor in activating gene regulation, and studies using ChIA-PET in human and mouse cells have found that many of these enhancers regulate the promoter distally [29], [30]. In the study of mouse Shh and SOX9 genes, it was also found that the enhancers interacting with the promoters of these genes were located outside the 1 Mb region [16], [31]. It is possible that the precise regulation of GRNs is partially because of the existence of these distant interactions that have shaped them into a complex regulatory conformation.
In 2009, Hi-C [32] and ChIA-PET [33] technologies were developed, which represented a milestone in the field. Hi-C [32] and ChIA-PET [33] technologies can uncover the spatial chromatin conformation, from which researchers can directly infer the interactions among genes and regulatory elements. Following that, various 3 C-based technologies, such as in situ Hi-C [34], single-cell Hi-C [35], Capture-Hi-C [36], and others were developed. These techniques have expanded the knowledge of GRN construction and illustrated the targeted relationships among them. Hi-C and other technologies have revealed the existence of TADs (Topologically Associating Domains). TADs are DNA fragments ranging from hundreds of kb to several Mb in the genome, containing one or more genes and regulatory elements, which can be regarded as a highly self-interacting genomic unit. Hi-C and other technologies have revealed the existence of TADs, which are thought to reflect the presence of physical interactions between DNA domains that organize chromosomes into distinct structural and functional units [37], [38]. TADs are highly conserved in different cell types and developmental stages; their positions are almost unchanged [39], [40], [41]. The relationship between TADs and gene transcription has always been a controversial? topic. During the differentiation of embryonic stem cells, the number of TADs is constantly changing [42]. The occurrence of some diseases or the body’s emergency response will also be accompanied by changes in TADs [43], [44], [45]. TAD-relative gene expression was also found to change when TAD borders were lost [46]. These examples demonstrate that TADs promote enhancer-promoter interaction in a cell type-independent manner in transcriptional regulation [38]. In embryonic development and cell differentiation, TADs become more structured, possibly because they instruct enhancer-promoter interaction and directly participate in transcriptional regulation [47].
In the matrix from Hi-C or other technical analysis, promoter-enhancer interactions were detected at high frequencies on the baseline of the domain formed by loop extrusion [48], [49]. These interactions form a certain chromatin structure through loop extrusion, which plays an important role in the interaction between promoters and enhancers in GRNs [1], [50]. Loop extrusion also promotes segregation, enhancer-promoter interactions, and other genomic functions [14]. It is important to dissect the constitution and biological effects of GRNs. The interaction between regulatory elements mediated by loop extrusion is generally more pronounced at the boundary of TADs and is generally limited within TADs. So loop extrusion regulates gene expression by promoting communication between enhancers and promoters, thereby affecting biological traits [14], [51], [52], [53], [54]. Additionally, insulators (boundary elements) are also critical to gene expression. Insulators are generally at the boundary of TADs, defined by CTCF and cohesin complexes [34], [37], [55], [56]. Because the insulator determines the boundary region, it determines which interactions in the genome can be formed [57], [58], [59]. For example, the deletion of the TAD boundary element of the Shh gene resulted in a weakened interaction between its promoter and enhancer, and its gene expression level also decreased by 50% [60]. The above series of studies have shown that in the genome, the regulatory elements are annotated by regulatory element identification technology, and the enhancer-promoter interactions are identified by chromatin conformation capture technology, represented by Hi-C. In each TAD, transcription factors mediate enhancer-promoter interactions to form spatial DNA loops, and there are also connections between multiple different DNA loops, forming a complex GRN to jointly regulate biological traits.
Overall, while the three-dimensional structure of chromatin is important for understanding gene regulatory networks, current research is mainly focused on describing the 3D genomic architecture, with less emphasis on its functional implications. However, the 3D structure provides important insights into gene regulatory activity and can serve as a starting point for further investigations into the function of GRNs. At the same time, most of these studies are carried out in population cells, so in the future, we should study the dynamics among different levels of chromatin structure and gene expression regulation in single cells and also combine epigenetics, phase separation, and other fields [14], [61], which will help us better understand the relationship between the 3D chromatin structure and GRN function.
3. Imaging technology supports the construction of gene regulatory networks
In recent years, rapid advances in imaging technology have provided an intuitive way to catalog gene regulatory networks. Visualizing the structure of chromatin and transcription complexes is a key step to understanding the transcription process and dissecting the biological effects of GRNs. The development of microscopy imaging technology allows us to observe the chromatin scaffold and the internal structure of chromatin [62], [63], [64], which is sometimes consistent with the results of Hi-C. For example, chromEM can integrate electron diffraction (ED) and electron tomography (ET) and visualize chromosome fibers and chromosome scaffolds [65] using modified soybean ascorbate peroxidase (APEX2) as a marker to observe the localization of chromatin in cells [66], [67]. Electron microscopy and confocal microscopy revealed highly and homogeneously condensed regions in chromatin and revealed that gene transcription occurred in the nuclear region [68], [69]. With increasing resolution, the proximity of loop anchors during gene transcription was observed under a high-resolution microscope [70], [71]. This series of studies shows that the electron microscope has been able to observe the real existence of transcriptional regulatory networks and reveal the impact of these spatial structures on gene expression. These have improved our understanding of the intuitive cooperation of GRNs and how they function during different biological processes.
Imaging technology is a powerful tool to investigate the process of transcriptional regulation. The physical distance of the site in the genome could be inferred to determine whether there is interaction. In the field of optical microscopy, DNA-FISH can judge the interaction by verifying the spatial distance between the two sites. Although this distance is related to the set threshold and resolution, it can be close to the real situation to some extent to study the interaction in gene transcription [72], [73], [74]. When DNA-FISH is combined with frozen sections, the genome can be observed at a higher resolution, such as in various cells where more than 20% of the chromosomal regions of the genome interact with other regions. This interaction is more pronounced in transcriptionally active regions [75], [76]. This phenomenon also suggests the universality and importance of regulatory networks throughout the genome. Chromatin folding and gene regulation are dynamic processes that require us to observe them in living cells. Genome editing techniques (such as the improved dCas9 system) can track the dynamic changes of the interaction between two genomic sites in living cells [77], [78], [79]. Although these techniques have shown the existence of GRN conformations, there has been no in-depth study of specific molecular mechanisms.
Using cryo-electron microscopy, the Xu Lab [80], [81] discovered the critical role of nucleosomes and the preinitiation complex (PIC) in the transcription process. For example, they obtained high-resolution nucleosomes and a human 26-subunit mediator cryomicroscope structure and revealed the mechanism by which the mediator and PIC are involved in the transcription process (Fig. 2). Previous studies have shown that nucleosomes play an important role in modifying proteins, processing RNA, and assembling transcription initiation complexes, but the mechanism has not been elucidated [82], [83]. In a recent study, Chen et al [84]. used cryo-electron microscopy to find that nucleosomes promote the organization of PIC-mediators on promoters by binding to the TFIIH subunit p52 and mediator subunits MED19 and MED26, which may facilitate transcriptional initiation. PIC-mediator exhibits multiple nucleosomes binding patterns, supporting the structural role of nucleosomes in PIC-mediator assembly coordination. This reveals the molecular mechanism of PIC-mediator organization on chromatin and emphasizes the importance of nucleosomes in regulating transcriptional initiation. A series of cryo-electron microscopy-based results advances our understanding of the mechanism of transcriptional regulation on the basis of the space network.
Fig. 2.
The pattern of mediator and PIC involvement in the transcription process.
Using microscopic imaging techniques, scientists have revealed the relationship between structure and gene expression regulation from another perspective. These results can be compared and integrated with the results of sequence-based techniques. From observing the interaction of regulatory elements with fluorescent groups to observing the high-resolution structure of transcriptional participants by cryo-electron microscopy, microscopic imaging technology has been called an indispensable part of the study of GRN. However, the quality and resolution of sample products can directly affect the results of observation. In addition, the throughput of these technologies is limited, and observations of transcription processes or participants are limited to a few critical biological processes. The improvement of sample preparation and resolution and the comprehensive tracking of transcription processes in living cells are the future development directions for understanding gene regulatory networks through imaging techniques.
4. Bioinformatics approaches to profiling gene regulatory networks
In the past two decades, advances in high-throughput biological data collection have greatly advanced our understanding of GRNs, which is based on bioinformatics analysis via mathematical modeling. In addition to the experimental methods described above for detecting chromatin contacts and GRN, network modeling techniques, such as network topology, transcriptome data modeling, chromatin conformation modeling, and deep learning, have been used to analyze its corresponding function and predict its behavior. GRN is often decomposed into basic building blocks with assumed functions, which extend the concept of structure-function relationships to the network schema level [85]. These patterns include auto-regulation (one gene regulates its own expression) [86], Feed Forward Loop (FFL) (one regulator regulates another regulator and a ternary structure of genes, which occurs more frequently in yeast GRN than other types[87]), Feedback Loops (looping action between multiple regulators and genes) [88], Single-Input Module (single regulators regulate multiple genes) and Dense Overlapping Regulon (composed of a set of regulators that control a set of output genes) [89] (Fig. 3). For functional studies, modularization by clustering can effectively identify co-expressed genes. The DREAM project uses genome-wide association study (GWAS) data to test and compare a module identification approach to human protein-protein interaction (PPI) data [88]. It has the advantage of using the diffusion state distances of the kernel clustering instead of the shortest path metric typically used to use PPI structure prediction functions to more accurately identify biological functions (Fig. 4A).
Fig. 3.
The topology of the gene regulatory network. Note: The dark-green dots indicate regulators, and the light-green dots represent genes.
Fig. 4.
Different visualizations of gene regulatory networks. (A) The protein-protein-interaction occurs during fibroblast differentiation into myoblasts. (B) Pol II ChIA-PET cis and trans interactions in mouse ESCs.
(A) (data source: STRING); (B) (data source: GSM1084137).
Combined with RNA-seq, machine learning algorithms are also an important technology for studying GRNs. GeneXPress first completes the mapping of GRNs based on a regression algorithm [90]. Regression-based methods use linear and nonlinear regression to predict mRNA levels as a function of chromatin markers and/or TF occupancy and can infer mRNA prediction models for individual conditions. GENIE3 [91], GRNBoost [115], LiPLike [93], and TIGRESS [92] are all based on this algorithm (Table 1). TIGRESS performs a series of feature selections for each target gene and uses randomization-based techniques to score evidence of regulatory interactions [92]. S. Feizi et al. used network deconvolution to solve the problem of identifying the direct relationship between variables connected in the network, which is a frequent problem in GRN mapping. Based on Bayesian inference [94], LeMoNe uses a centroid clustering method to infer GRNs, assigning genes and conditions to modules, and then assigning regulatory programs to gene sets [114]. ARACNe-AP applied partial correlation coefficients and data processing inequalities to distinguish direct and indirect dependencies, relating changes in TF mRNA levels to sets of other genes’ mRNA levels [95]. These methods infer cellular networks mainly using regression, correlation analysis, and Bayesian networks as the main tools, which require the least input data and have wide applicability. However, due to the high experimental cost and the fact that the sample size is much smaller than the number of TFs, the results are susceptible to noise. Various machine learning approaches have performed poorly on computationally simulated data evaluations of E. coli and Saccharomyces cerevisiae [96], [97]. This type of algorithm is also highly dependent on the quality of the data and the uncertainty in determining the relevant parameters. Noise, modeling form, parameter estimation methods, and the amount of experimental data are all crucial factors that affect the results [98]. Meanwhile, different analysis strategies and model choices can lead to varying effects of network inference. It is also a complex problem to choose the appropriate scheme to ensure the biological correctness of the results [99]. Therefore, in the process of exploring the GRN, we need the integrative analysis of multiple types of omics data, such as 3D genomic data, which opens a novel path.
Table 1.
Summary of tools used for GRN analyze.
Tools | Description | Data | Performance | Restriction | Link | Ref |
---|---|---|---|---|---|---|
GeneXPress | regression algorithm | bulk transcriptomics | Accurately predict regulator function, targets and the experimental conditions under this regulation occurs | 1. Fail to identify certain regulatory relationships and may occasionally predict regulatory relationships that do not hold; 2. A gene may belong only to a single module |
(https://pypi.org/project/GeneXpress/#files) | [90] |
TIGRESS | regression algorithm | bulk transcriptomics | Expresses the GRN inference problem as a feature selection problem, and solves it with the popular LARS feature selection method combined with stability selection | Heavy formatting requirements | (http://cbio.ensmp.fr/tigress) | [92] |
LeMoNe | Bayesian inference | bulk transcriptomics | Learning process is considerably faster for larger data sets | -- | (http://bioinformatics.psb.ugent.be/software) | [114] |
Network Deconvolution | Network deconvolution | bulk transcriptomics | 1. Distinguishing direct targets in gene expression regulatory networks; 2. Recognizing directly interacting amino-acid residues for protein structure prediction from sequence alignments; 3. And distinguishing strong collaborations in co-authorship social networks using connectivity information alone |
Network deconvolution assumes that networks are linear time-invariant flow-preserving operators, which excludes nonlinear dynamic networks as well as hidden sinks or sources. | (http://compbio.mit.edu/nd/index.html) | [94] |
ARACNe-AP | Information theory | bulk transcriptomics | A dramatic improvement in computational performance (200 × on average) over the previous methodology, while preserving the Mutual Information estimator and the Network inference accuracy of the original algorithm | -- | (http://sourceforge.net/projects/aracne-ap) | [95] |
GENIE3 | regression algorithm | bulk transcriptomics | Work well in the presence of a large number of genes, are fast to compute, and are scalable | Prior knowledge about transcription factors can have a significant impact on the results | (https://bioconductor.org/packages/release/bioc/html/GENIE3.html) | [91] |
GRNBoost | regression algorithm | bulk transcriptomics | Implementation drastically reduces the time needed to infer a GRN | Statically performs workload scheduling | (http://arboreto.readthedocs.io) | [115] |
LiPLike | regression algorithm | bulk transcriptomics | Higher accuracy and removing false positive identifications from GRN predictions | -- | (https://gitlab.com/Gustafsson-lab/liplike) | [93] |
BART3D | genomic differential chromatin interaction | Hi-C | Can infer TRs associated with genome-wide differential chromatin interactions | Genomic distance needs to be set manually | (https://github.com/zanglab/bart3d) | [100] |
3CPET | non-parametric Bayesian | ChIA-PET | Infer the most likely set of protein networks likely to be involved in maintaining chromatin interactions with reasonable performance and good stability | memory and time requirements for 3CPET gradually increase with the DNA–DNA corpus size | (http://www.bioconductor.org/packages/release/bioc/html/R3CPET.html) | [103] |
By integrating 3D genome data with different ∼omics datasets, not only the identification of GRNs, but also the precise location of interactions between enhancer-promoter and other transcriptional regulatory elements can be achieved (Fig. 4B). For example, based on Hi-C and ChIP-seq data, BART3D designed a DCI scoring algorithm to infer TRs associated with genome-wide differential chromatin interactions and validated that genomic regions containing CTCF or RAD21 binding sites exhibited reduced levels of chromatin interactions after knockout [100]. Ma et al. [101] used single-cell Hi-C data to demonstrate that within differentiated and undifferentiated cells, the association of TFs in spatial networks is closely related to GRNs, and that lineage-specific TF binding and chromosomal subcompartment segregation are closely related potential causal relationships [101]. By modeling Hi-C data, Johan et al. found a characteristic of chromosomes in interphase during mitosis. Especially at anaphase, chromosomes fold into contiguous arrays of rings condensed around a central axis. The chromatin loops are nested within a ∼400-kb outer loop separated by an ∼80-kb inner loop [102]. This is an important complement to the observations made by electron microscopy. 3 CPET is based on a nonparametric Bayesian approach that infers the most likely protein complexes involved in maintaining chromatin interactions and the regions they may control [103]. Based on 3D genome technology combined with gene expression and chromatin state data, the GRN-loop model of induced early neural lineage commitment effectively identifies driving transcription factors and enhancers and then validates the impact of predicted genetic perturbations on lineage commitment [97]. We can more accurately interpret enhancer-promoter interactions by applying machine learning algorithms, epigenetic data, and three-dimensional genome technology for mutual verification. RIPPLE combines the published chromosome conformation capture (3 C) dataset with a minimal regulatory genome dataset to construct a minimal classifier by combining random forest and group lasso-based multi-task learning to predict enhancer-priming in cells and subinteractions in a circuit-specific manner [104]. IM-PET is based on identifying discriminative features of EP pairs, including enhancer and target promoter activity profile correlations, transcription factor and target promoter correlations, the co-evolution of enhancer and target promoter, and the interactions between enhancer and target promoter. The distance constraints were trained on genome sequence, three histone modification ChIP-Seq and RNA-Seq/microarray data, judged by the random forest method, and validated by ChIA-PET [105]. In addition, accurate identification of the target genes of regulatory elements is very important for the study of enhancer-promoter interactions. Sean Whalen et al. developed an algorithm that can detect single enhancer-promoter interactions in multiple cell lines with an error rate 15 times lower than simply using the nearest target gene [106]. HiC-Reg combines published Hi-C datasets with one-dimensional regulatory genome datasets (such as chromatin marks, structural and transcription factor proteins, and chromatin accessibility) to predict two interaction counts between genomic loci. This approach breaks through the traditional binary classification of interacting vs. non-interacting and can predict interactions for new chromosomes or cell lines of interest [107].
Deep learning is another cutting-edge weapon for studying GRNs. It integrates multi-omics data for analysis and prediction. Simultaneously, biological function analysis is also carried out based on identifying GRNs. These tools identify transcriptional regulatory elements by analyzing sequence features or further differentiate network differences in tissue characteristics and cell types based on the identification of transcriptional regulatory networks. DeepSEA was developed as a fully sequence-based algorithmic framework for noncoding-variant effect prediction. It learns regulatory sequence codes from genomic sequences by simultaneously predicting large-scale chromatin-profiling data, including TF binding, DHS, and histone mark profiles [108]. BiRen uses a deep learning hybrid architecture to predict enhancers based on DNA sequence alone. BiRen exhibited superior accuracy, robustness, and generalizability in enhancer prediction relative to other state-of-the-art enhancer predictors based on sequence characteristics [109]. DGRNS adopts a hybrid deep learning framework. With the help of recurrent neural networks and convolutional neural networks, not only statistical features but also time-related features can be extracted. Based on single-cell RNA-Seq technology, the regulatory network was classified and analyzed to explore network differences between biological tissues [110]. At the same time, deep learning is also applied to the prediction of three-dimensional genome structures. It is also helpful for the identification of GRNs. Akita converts input DNA sequences into predicted locus-specific genome folds based on convolutional neural network, Hi-C, and Micro-C data and then estimates entire contact maps [111]. Similar to Akita, DeepC also predicts genome folding through the convolutional neural network, but differs in data preprocessing and training methods [112]. DeepC employs a non-linear normalization objective that requires pre-training models on a large number of epigenomic profiles and transferring weights to their full models. The calculated results of the differences between the two methods merit further discussion of their advantages and disadvantages in future comparative studies.
Other techniques include methods such as protein interactomes, eQTL data, the hierarchy of tissues, etc., which are jointly applied to the prediction of GRNs. Recent advances in genomic technology and computational modeling have revolutionized our ability to model GRNs. Researchers have applied reverse engineering methods to construct GRNs in cancer cells to improve the accuracy of the comparison of the differences in GRN between normal and cancer cells [113]. In recent years, the interpretation of multilayer networks has been a key issue for a comprehensive understanding of gene regulatory systems. Although several statistical and machine-learning approaches have been developed and applied to uncover mechanisms, a comprehensive understanding of large-scale multilayer networks remains a challenge. Data-driven methods based on artificial intelligence and multi-omics data, especially 3D genomics, provide biologically reliable results and are the direction of bioinformatics efforts in this field.
5. Summary and outlook
Research on GRNs is an important subject in the field of precision biology. Although the continuous development of chromatin conformation capture technology provides good technical support for the interpretation of GRNs, there is still much room for improvement. Experimentally, the quality of the data produced by different laboratories is uneven, with poor repeatability, and the data also has low resolution, a poor signal-to-noise ratio, and other issues. In terms of data processing, there are differences in the evaluation criteria of different software, resulting in differences in the results of even the same data, such as the number and overlap of loops obtained by Homer, HiCCUPs and FitHic [116] when loop calling. Homer and FitHic directly model individual interacting elements, considering the effects of systematic bias or linear distance across the genome. Among them, FitHic estimates that the interaction of random polymers is affected by distance by modeling the relationship between interaction and space twice and using the binomial distribution to model the interaction, giving the significance of the interaction. While HiCCUPS uses the Poisson process to consider the effects of neighbor interactions or other hierarchies, the first type of algorithm can identify more significant interactions from the results, but the false discovery rate is relatively high. While the latter is computationally more rigorous, it will lose some potentially practical information. More importantly, the conclusions of different studies vary greatly. For example, there are many inconsistent results in TADs and more studies are needed on the effects of chromatin structure on biological functions. Changes in TAD were found to cause changes in gene expression in a study of the mouse α-globin gene [117]. In mouse cells, silencing of the male X chromosome also leads to the rearrangement of the associated TAD [118]. These all illustrate the important influence of TADs on gene expression and traits. However, in other studies, it was found that the destruction of TADs structure had little effect on gene expression [16]. Alternatively, when it comes to the presence or absence of TAD, ChromEMT produces inconsistent results [65] when compared to the traditional concept. In addition, although MERFISH imaging technology observes TAD-like domains at the single-cell level, its high heterogeneity contradicts the high conservation of TAD in the traditional concept [119]. At the same time, the need for a unified TAD identification standard also makes the credibility of the results questionable. The inferred results of different TAD identification methods are also different. The resolution size is another factor that affects the results [120]. Current criteria for evaluating TAD identification tools include robustness concerning bin size and normalization strategy, the minimum required data volume, reproducibility, biological relevance, and computational efficiency [121]. As biological researchers, however, we should pay more attention to interpreting TAD for its biological meaning itself than for the statistical results of computational observations. This needs to be verified by a more suitable data set and other corresponding experimental observation methods. These results make us think about whether this is a defect of electron microscopy or whether TAD is only a statistical model; so far, there is no reasonable explanation.
At present, the research on gene regulatory interactions has focused on how enhancers interact with promoters under the mediation of TF. In the human genome, less than 15% of genes and regulatory elements have been annotated [122], and the function of the remaining 85% of non-coding sequences is unknown. Recently, the function of silencing regulatory elements in gene regulation has been reported [123], [124], [125]. However, because these reports only covered a portion of the silencer's genome region, whole-genome silencer identification is an important supplement to GRN. Additionally, with the dynamic changes in organisms, the corresponding GRNs will also change. Most of the current technologies show static chromatin conformation and regulatory networks. In more cases, we only describe that changes in the spatial conformation will affect gene transcriptional regulation or that changes in gene expression will cause changes in chromatin conformation. Can we change chromatin conformation, gene expression, or even phenotype through these regularities? A recent study reported that cell identity can be changed by introducing exogenous enhancer elements into cells [126], which makes us realize the substantial regulation of gene expression by spatial conformation. The development of various fields has promoted the completion of GRNs.
We briefly summarize the relationship between chromatin spatial structure and gene expression regulation. The chromatin spatial structure and epigenetic program play pivotal roles in the GRN. The chromatin spatial structure embraces loops and TADs. The TADs include vast cis-loops. The loop extrusion underlies the TAD formation [127]. The loops formed by the proximity of enhancer-promoter or silencer-promoter, which are mediated by TFs. These will then get involved in gene expression regulation. Furthermore, the boundary elements are typically enriched in the TAD boundaries, which frequently bind the CTCF. And repression of the boundary elements could alter the structure of TADs, which would have a further impact on gene expression (Fig. 5).
Fig. 5.
The relationship between chromatin spatial structure and gene regulatory network.
Lastly, we realize the high-throughput, multi-scale, and multi-modal imaging technology will provide an integrated view of chromatin organization and structural biology in its native genome context. By integrating multiple-omics datasets, a good bioinformatics algorithm can identify relationships among genes and regulatory elements, which would greatly facilitate the unveiling of the comprehensive map of individual cell GRNs, thus allowing a novel and fundamental understanding of how a cell operates and initiates life. In the near future, through the cooperation of imaging technology, genomics technology, computer science, physics, and other fields, a complete, dynamic, and precise GNR map will be achieved, which will facilitate the gene function analysis for the purpose.
CRediT authorship contribution statement
Xiusheng Zhu: Conceptualization, original draft, Writing – review & editing, Methodology, Formal analysis. Qitong Huang: Writing – original draft, Conceptualization, Formal analysis, Writing – review & editing. Jing Luo: Writing – original draft, Conceptualization, Formal analysis, Writing – review & editing. Dashuai Kong: Writing – review & editing, Conceptualization. Yubo Zhang: Conceptualization, Writing – original draft, Writing – review & editing, Supervision.
Declaration of Competing Interest
The authors declare that there are no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
Acknowledgments
This work was supported by the National Key Research and Development Program of China (No. 2018YFA0903201); the Agricultural Science and Technology Innovation Program; the National Natural Science Foundation of China (No. 31970592, 32002173 and 32202653); the Guangdong Basic and Applied Basic Research Foundation (No. 2022A1515010766); the Shenzhen Science and Technology Program (Grant No. KCXFZ20201221173205015 and RCBS20210609104512021).
References
- 1.Dekker J., Mirny L. The 3D genome as moderator of chromosomal communication. Cell. 2016;164:1110–1121. doi: 10.1016/j.cell.2016.02.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Preissl S., Gaulton K.J., Ren B. Characterizing cis-regulatory elements using single-cell epigenomics. Nat Rev Genet. 2022 doi: 10.1038/s41576-022-00509-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Jacob F., Monod J. Genetic regulatory mechanisms in the synthesis of proteins. J Mol Biol. 1961;3:318–356. doi: 10.1016/s0022-2836(61)80072-7. [DOI] [PubMed] [Google Scholar]
- 4.Rinn J.L., Chang H.Y. Genome regulation by long noncoding RNAs. Annu Rev Biochem. 2012;81:145–166. doi: 10.1146/annurev-biochem-051410-092902. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Stadhouders R., Filion G.J., Graf T. Transcription factors and 3D genome conformation in cell-fate decisions. Nature. 2019;569:345–354. doi: 10.1038/s41586-019-1182-7. [DOI] [PubMed] [Google Scholar]
- 6.Arnold C.D., Gerlach D., Stelzer C., Boryń Ł.M., Rath M., Stark A. Genome-wide quantitative enhancer activity maps identified by STARR-seq. Sci (N Y, NY) 2013;339:1074–1077. doi: 10.1126/science.1232542. [DOI] [PubMed] [Google Scholar]
- 7.Andersson R., Sandelin A., Danko C.G. A unified architecture of transcriptional regulatory elements. Trends Genet: TIG. 2015;31:426–433. doi: 10.1016/j.tig.2015.05.007. [DOI] [PubMed] [Google Scholar]
- 8.Heinz S., Romanoski C.E., Benner C., Glass C.K. The selection and function of cell type-specific enhancers. Nat Rev Mol Cell Biol. 2015;16:144–154. doi: 10.1038/nrm3949. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Soshnev A.A., Josefowicz S.Z., Allis C.D. Greater than the sum of parts: complexity of the dynamic epigenome. Mol Cell. 2016;62:681–694. doi: 10.1016/j.molcel.2016.05.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Pope S.D., Medzhitov R. Emerging principles of gene expression programs and their regulation. Mol Cell. 2018;71:389–397. doi: 10.1016/j.molcel.2018.07.017. [DOI] [PubMed] [Google Scholar]
- 11.Sun D.Q., Tian L., Ma B.G. Spatial organization of the transcriptional regulatory network of Saccharomyces cerevisiae. FEBS Lett. 2019;593:876–884. doi: 10.1002/1873-3468.13371. [DOI] [PubMed] [Google Scholar]
- 12.Lee T.I., Young R.A. Transcriptional regulation and its misregulation in disease. Cell. 2013;152:1237–1251. doi: 10.1016/j.cell.2013.02.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Levine M., Cattoglio C., Tjian R. Looping back to leap forward: transcription enters a new era. Cell. 2014;157:13–25. doi: 10.1016/j.cell.2014.02.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Oudelaar A.M., Higgs D.R. The relationship between genome structure and function. Nat Rev Genet. 2021;22:154–168. doi: 10.1038/s41576-020-00303-x. [DOI] [PubMed] [Google Scholar]
- 15.Lupiáñez D.G., Kraft K., Heinrich V., Krawitz P., Brancati F., Klopocki E., et al. Disruptions of topological chromatin domains cause pathogenic rewiring of gene-enhancer interactions. Cell. 2015;161:1012–1025. doi: 10.1016/j.cell.2015.04.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Franke M., Ibrahim D.M., Andrey G., Schwarzer W., Heinrich V., Schöpflin R., et al. Formation of new chromatin domains determines pathogenicity of genomic duplications. Nature. 2016;538:265–269. doi: 10.1038/nature19800. [DOI] [PubMed] [Google Scholar]
- 17.Banerji J., Rusconi S., Schaffner W. Expression of a beta-globin gene is enhanced by remote SV40 DNA sequences. Cell. 1981;27:299–308. doi: 10.1016/0092-8674(81)90413-x. [DOI] [PubMed] [Google Scholar]
- 18.Rada-Iglesias A., Bajpai R., Swigut T., Brugmann S.A., Flynn R.A., Wysocka J. A unique chromatin signature uncovers early developmental enhancers in humans. Nature. 2011;470:279–283. doi: 10.1038/nature09692. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Pradeepa M.M., Grimes G.R., Kumar Y., Olley G., Taylor G.C., Schneider R., et al. Histone H3 globular domain acetylation identifies a new class of enhancers. Nat Genet. 2016;48:681–686. doi: 10.1038/ng.3550. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Song L., Crawford G.E. DNase-seq: a high-resolution technique for mapping active gene regulatory elements across the genome from mammalian cells. Cold Spring Harbor protocols. 2010;2010:pdb.prot5384. https://doi.org/10.1101/pdb.prot5384. [DOI] [PMC free article] [PubMed]
- 21.Buenrostro J.D., Giresi P.G., Zaba L.C., Chang H.Y., Greenleaf W.J. Transposition of native chromatin for fast and sensitive epigenomic profiling of open chromatin, DNA-binding proteins and nucleosome position. Nat Methods. 2013;10:1213–1218. doi: 10.1038/nmeth.2688. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Whyte W.A., Orlando D.A., Hnisz D., Abraham B.J., Lin C.Y., Kagey M.H., et al. Master transcription factors and mediator establish super-enhancers at key cell identity genes. Cell. 2013;153:307–319. doi: 10.1016/j.cell.2013.03.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Marchal C., Singh N., Batz Z., Advani J., Jaeger C., Corso-Díaz X., et al. High-resolution genome topology of human retina uncovers super enhancer-promoter interactions at tissue-specific and multifactorial disease loci. Nat Commun. 2022;13:5827. doi: 10.1038/s41467-022-33427-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Long H.K., Prescott S.L., Wysocka J. Ever-changing landscapes: transcriptional enhancers in development and evolution. Cell. 2016;167:1170–1187. doi: 10.1016/j.cell.2016.09.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Hay D., Hughes J.R., Babbs C., Davies J.O.J., Graham B.J., Hanssen L., et al. Genetic dissection of the α-globin super-enhancer in vivo. Nat Genet. 2016;48:895–903. doi: 10.1038/ng.3605. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Zhu X., Huang L., Li Q., Zhang Y. Start from scratch: precisely identify massive active enhancers by sequencing. BioRxiv. 2019 doi: 10.1101/604686. [DOI] [Google Scholar]
- 27.Will A.J., Cova G., Osterwalder M., Chan W.L., Wittler L., Brieske N., et al. Composition and dosage of a multipartite enhancer cluster control developmental expression of Ihh (Indian hedgehog) Nat Genet. 2017;49:1539–1545. doi: 10.1038/ng.3939. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Xie S., Duan J., Li B., Zhou P., Hon G.C. Multiplexed engineering and analysis of combinatorial enhancer activity in single cells. Mol Cell. 2017;66(285–99) doi: 10.1016/j.molcel.2017.03.007. [DOI] [PubMed] [Google Scholar]
- 29.Sanyal A., Lajoie B.R., Jain G., Dekker J. The long-range interaction landscape of gene promoters. Nature. 2012;489:109–113. doi: 10.1038/nature11279. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Zhang Y., Wong C.H., Birnbaum R.Y., Li G., Favaro R., Ngan C.Y., et al. Chromatin connectivity maps reveal dynamic promoter-enhancer long-range associations. Nature. 2013;504:306–310. doi: 10.1038/nature12716. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Sagai T., Hosoya M., Mizushina Y., Tamura M., Shiroishi T. Elimination of a long-range cis-regulatory module causes complete loss of limb-specific Shh expression and truncation of the mouse limb. Dev (Camb, Engl) 2005;132:797–803. doi: 10.1242/dev.01613. [DOI] [PubMed] [Google Scholar]
- 32.Fullwood M.J., Liu M.H., Pan Y.F., Liu J., Xu H., Mohamed Y.B., et al. An oestrogen-receptor-alpha-bound human chromatin interactome. Nature. 2009;462:58–64. doi: 10.1038/nature08497. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Lieberman-Aiden E., van Berkum N.L., Williams L., Imakaev M., Ragoczy T., Telling A., et al. Comprehensive mapping of long-range interactions reveals folding principles of the human genome. Sci (N Y, NY) 2009;326:289–293. doi: 10.1126/science.1181369. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Rao S.S., Huntley M.H., Durand N.C., Stamenova E.K., Bochkov I.D., Robinson J.T., et al. A 3D map of the human genome at kilobase resolution reveals principles of chromatin looping. Cell. 2014;159:1665–1680. doi: 10.1016/j.cell.2014.11.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Nagano T., Lubling Y., Stevens T.J., Schoenfelder S., Yaffe E., Dean W., et al. Single-cell Hi-C reveals cell-to-cell variability in chromosome structure. Nature. 2013;502:59–64. doi: 10.1038/nature12593. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Schoenfelder S., Sexton T., Chakalova L., Cope N.F., Horton A., Andrews S., et al. Preferential associations between co-regulated genes reveal a transcriptional interactome in erythroid cells. Nat Genet. 2010;42:53–61. doi: 10.1038/ng.496. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Dixon J.R., Selvaraj S., Yue F., Kim A., Li Y., Shen Y., et al. Topological domains in mammalian genomes identified by analysis of chromatin interactions. Nature. 2012;485:376–380. doi: 10.1038/nature11082. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Nora E.P., Lajoie B.R., Schulz E.G., Giorgetti L., Okamoto I., Servant N., et al. Spatial partitioning of the regulatory landscape of the X-inactivation centre. Nature. 2012;485:381–385. doi: 10.1038/nature11049. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Battulin N., Fishman V.S., Mazur A.M., Pomaznoy M., Khabarova A.A., Afonnikov D.A., et al. Comparison of the three-dimensional organization of sperm and fibroblast genomes using the Hi-C approach. Genome Biol. 2015;16:77. doi: 10.1186/s13059-015-0642-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Dekker J., Heard E. Structural and functional diversity of Topologically Associating Domains. FEBS Lett. 2015;589:2877–2884. doi: 10.1016/j.febslet.2015.08.044. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Dixon J.R., Gorkin D.U., Ren B. Chromatin domains: the unit of chromosome organization. Mol Cell. 2016;62:668–680. doi: 10.1016/j.molcel.2016.05.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Ke Y., Xu Y., Chen X., Feng S., Liu Z., Sun Y., et al. 3D chromatin structures of mature gametes and structural reprogramming during mammalian embryogenesis. Cell. 2017;170(367–81) doi: 10.1016/j.cell.2017.06.029. [DOI] [PubMed] [Google Scholar]
- 43.Li W., Gong K., Li Q., Alber F., Zhou X.J. Hi-corrector: a fast, scalable and memory-efficient package for normalizing large-scale Hi-C data. Bioinforma (Oxf, Engl) 2015;31:960–962. doi: 10.1093/bioinformatics/btu747. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Hnisz D., Weintraub A.S., Day D.S., Valton A.L., Bak R.O., Li C.H., et al. Activation of proto-oncogenes by disruption of chromosome neighborhoods. Sci (N Y, NY) 2016;351:1454–1458. doi: 10.1126/science.aad9024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Lupiáñez D.G., Spielmann M., Mundlos S. Breaking TADs: how alterations of chromatin domains result in disease. Trends Genet TIG. 2016;32:225–237. doi: 10.1016/j.tig.2016.01.003. [DOI] [PubMed] [Google Scholar]
- 46.Chathoth K.T., Mikheeva L.A., Crevel G., Wolfe J.C., Hunter I., Beckett-Doyle S., et al. The role of insulators and transcription in 3D chromatin organization of flies. Genome Res. 2022;32:682–698. doi: 10.1101/gr.275809.121. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Bonev B., Cavalli G. Organization and function of the 3D genome. Nat Rev Genet. 2016;17:661–678. doi: 10.1038/nrg.2016.112. [DOI] [PubMed] [Google Scholar]
- 48.Vian L., Pękowska A., Rao S.S.P., Kieffer-Kwon K.R., Jung S., Baranello L., et al. The energetics and physiological impact of cohesin extrusion. Cell. 2018;173(1165–78) doi: 10.1016/j.cell.2018.03.072. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Barrington C., Georgopoulou D., Pezic D., Varsally W., Herrero J., Hadjur S. Enhancer accessibility and CTCF occupancy underlie asymmetric TAD architecture and cell type specific genome topology. Nat Commun. 2019;10:2908. doi: 10.1038/s41467-019-10725-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Chang L.H., Ghosh S., Noordermeer D. TADs and their borders: free movement or building a wall? J Mol Biol. 2020;432:643–652. doi: 10.1016/j.jmb.2019.11.025. [DOI] [PubMed] [Google Scholar]
- 51.Rowley M.J., Corces V.G. Organizational principles of 3D genome architecture. Nat Rev Genet. 2018;19:789–800. doi: 10.1038/s41576-018-0060-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Banigan E.J., Mirny L.A. Loop extrusion: theory meets single-molecule experiments. Curr Opin Cell Biol. 2020;64:124–138. doi: 10.1016/j.ceb.2020.04.011. [DOI] [PubMed] [Google Scholar]
- 53.Davidson I.F., Peters J.M. Genome folding through loop extrusion by SMC complexes. Nat Rev Mol Cell Biol. 2021;22:445–464. doi: 10.1038/s41580-021-00349-7. [DOI] [PubMed] [Google Scholar]
- 54.Peters J.M. How DNA loop extrusion mediated by cohesin enables V(D)J recombination. Curr Opin Cell Biol. 2021;70:75–83. doi: 10.1016/j.ceb.2020.11.007. [DOI] [PubMed] [Google Scholar]
- 55.Parelho V., Hadjur S., Spivakov M., Leleu M., Sauer S., Gregson H.C., et al. Cohesins functionally associate with CTCF on mammalian chromosome arms. Cell. 2008;132:422–433. doi: 10.1016/j.cell.2008.01.011. [DOI] [PubMed] [Google Scholar]
- 56.Wendt K.S., Yoshida K., Itoh T., Bando M., Koch B., Schirghuber E., et al. Cohesin mediates transcriptional insulation by CCCTC-binding factor. Nature. 2008;451:796–801. doi: 10.1038/nature06634. [DOI] [PubMed] [Google Scholar]
- 57.de Wit E., Vos E.S., Holwerda S.J., Valdes-Quezada C., Verstegen M.J., Teunissen H., et al. CTCF binding polarity determines chromatin looping. Mol Cell. 2015;60:676–684. doi: 10.1016/j.molcel.2015.09.023. [DOI] [PubMed] [Google Scholar]
- 58.Guo Y., Xu Q., Canzio D., Shou J., Li J., Gorkin D.U., et al. CRISPR inversion of CTCF sites alters genome topology and enhancer/promoter function. Cell. 2015;162:900–910. doi: 10.1016/j.cell.2015.07.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Sanborn A.L., Rao S.S., Huang S.C., Durand N.C., Huntley M.H., Jewett A.I., et al. Chromatin extrusion explains key features of loop and domain formation in wild-type and engineered genomes. Proc Natl Acad Sci USA. 2015;112:E6456–E6465. doi: 10.1073/pnas.1518552112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Paliou C., Guckelberger P., Schöpflin R., Heinrich V., Esposito A., Chiariello A.M., et al. Preformed chromatin topology assists transcriptional robustness of Shh during limb development. Proc Natl Acad Sci USA. 2019;116:12390–12399. doi: 10.1073/pnas.1900672116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Kempfer R., Pombo A. Methods for mapping 3D chromosome architecture. Nat Rev Genet. 2020;21:207–226. doi: 10.1038/s41576-019-0195-2. [DOI] [PubMed] [Google Scholar]
- 62.Frank J. In: Electron Tomography. US. Frank J., editor. Springer International Publishing; 1992. Introduction: principles of electron tomography; pp. 1–13. [Google Scholar]
- 63.Masters B.R. In: Superresolution Optical Microscopy. Rhodes W.T., Adibi A., Asakura T., Hänsch T.W., Krausz F., Masters B.R., Venghaus H., Weber H., Weinfurter H., Midorikawa K., editors. Springer International Publishing; US: 2020. Abbe’s theory of image formation in the microscope; pp. 65–107. [Google Scholar]
- 64.Phengchat R., Malac M., Hayashida M. Chromosome inner structure investigation by electron tomography and electron diffraction in a transmission electron microscope. Chromosome Res: Int J Mol, Supramol Evolut Asp Chromosome Biol. 2021;29:63–80. doi: 10.1007/s10577-021-09661-6. [DOI] [PubMed] [Google Scholar]
- 65.Ou H.D., Phan S., Deerinck T.J., Thor A., Ellisman M.H., O'Shea C.C. ChromEMT: Visualizing 3D chromatin structure and compaction in interphase and mitotic cells. Sci (N Y, NY) 2017:357. doi: 10.1126/science.aag0025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Martell J.D., Deerinck T.J., Sancak Y., Poulos T.L., Mootha V.K., Sosinsky G.E., et al. Engineered ascorbate peroxidase as a genetically encoded reporter for electron microscopy. Nat Biotechnol. 2012;30:1143–1148. doi: 10.1038/nbt.2375. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Lam S.S., Martell J.D., Kamer K.J., Deerinck T.J., Ellisman M.H., Mootha V.K., et al. Directed evolution of APEX2 for electron microscopy and proximity labeling. Nat Methods. 2015;12:51–54. doi: 10.1038/nmeth.3179. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Monneron A., Bernhard W. Fine structural organization of the interphase nucleus in some mammalian cells. J Ultrastruct Res. 1969;27:266–288. doi: 10.1016/s0022-5320(69)80017-1. [DOI] [PubMed] [Google Scholar]
- 69.Verschure P.J., van Der Kraan I., Manders E.M., van Driel R. Spatial relationship between transcription sites and chromosome territories. J Cell Biol. 1999;147:13–24. doi: 10.1083/jcb.147.1.13. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Dundr M., Misteli T. Biogenesis of nuclear bodies. Cold Spring Harb Perspect Biol. 2010;2:a000711. doi: 10.1101/cshperspect.a000711. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Mao Y.S., Zhang B., Spector D.L. Biogenesis and function of nuclear bodies. Trends Genet: TIG. 2011;27:295–306. doi: 10.1016/j.tig.2011.05.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Barbieri M., Xie S.Q., Torlai Triglia E., Chiariello A.M., Bianco S., de Santiago I., et al. Active and poised promoter states drive folding of the extended HoxB locus in mouse embryonic stem cells. Nat Struct Mol Biol. 2017;24:515–524. doi: 10.1038/nsmb.3402. [DOI] [PubMed] [Google Scholar]
- 73.Barutcu A.R., Maass P.G., Lewandowski J.P., Weiner C.L., Rinn J.L.A. TAD boundary is preserved upon deletion of the CTCF-rich Firre locus. Nat Commun. 2018;9:1444. doi: 10.1038/s41467-018-03614-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Maass P.G., Barutcu A.R., Weiner C.L., Rinn J.L. Inter-chromosomal contact properties in live-cell imaging and in Hi-C. Mol Cell. 2018;69(1039–45) doi: 10.1016/j.molcel.2018.02.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Branco M.R., Pombo A. Intermingling of chromosome territories in interphase suggests role in translocations and transcription-dependent associations. PLoS Biol. 2006;4 doi: 10.1371/journal.pbio.0040138. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Rosin L.F., Nguyen S.C., Joyce E.F. Condensin II drives large-scale folding and spatial partitioning of interphase chromosomes in Drosophila nuclei. PLoS Genet. 2018;14 doi: 10.1371/journal.pgen.1007393. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Fu Y., Rocha P.P., Luo V.M., Raviram R., Deng Y., Mazzoni E.O., 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. doi: 10.1038/ncomms11707. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.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 doi: 10.1093/nar/gkw066. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Wang S., Su J.H., Zhang F., Zhuang X. An RNA-aptamer-based two-color CRISPR labeling system. Sci Rep. 2016;6:26857. doi: 10.1038/srep26857. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.He S., Wu Z., Tian Y., Yu Z., Yu J., Wang X., et al. Structure of nucleosome-bound human BAF complex. Sci (N Y, NY) 2020;367:875–881. doi: 10.1126/science.aaz9761. [DOI] [PubMed] [Google Scholar]
- 81.Zheng H., Qi Y., Hu S., Cao X., Xu C., Yin Z., et al. Identification of Integrator-PP2A complex (INTAC), an RNA polymerase II phosphatase. Sci (N Y, NY) 2020:370. doi: 10.1126/science.abb5872. [DOI] [PubMed] [Google Scholar]
- 82.Rhee H.S., Bataille A.R., Zhang L., Pugh B.F. Subnucleosomal structures and nucleosome asymmetry across a genome. Cell. 2014;159:1377–1388. doi: 10.1016/j.cell.2014.10.054. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Allen B.L., Taatjes D.J. The Mediator complex: a central integrator of transcription. Nat Rev Mol Cell Biol. 2015;16:155–166. doi: 10.1038/nrm3951. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Chen X., Wang X., Liu W., Ren Y., Qu X., Li J., et al. Structures of +1 nucleosome-bound PIC-Mediator complex. Sci (N Y, NY) 2022;378:62–68. doi: 10.1126/science.abn8131. [DOI] [PubMed] [Google Scholar]
- 85.Monteiro P.T., Pedreira T., Galocha M., Teixeira M.C., Chaouiya C. Assessing regulatory features of the current transcriptional network of Saccharomyces cerevisiae. Sci Rep. 2020;10:17744. doi: 10.1038/s41598-020-74043-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Rosenfeld N., Elowitz M.B., Alon U. Negative autoregulation speeds the response times of transcription networks. J Mol Biol. 2002;323:785–793. doi: 10.1016/s0022-2836(02)00994-4. [DOI] [PubMed] [Google Scholar]
- 87.Mangan S., Itzkovitz S., Zaslaver A., Alon U. The incoherent feed-forward loop accelerates the response-time of the gal system of Escherichia coli. J Mol Biol. 2006;356:1073–1081. doi: 10.1016/j.jmb.2005.12.003. [DOI] [PubMed] [Google Scholar]
- 88.Thieffry D. Dynamical roles of biological regulatory circuits. Brief Bioinforma. 2007;8:220–225. doi: 10.1093/bib/bbm028. [DOI] [PubMed] [Google Scholar]
- 89.Shen-Orr S.S., Milo R., Mangan S., Alon U. Network motifs in the transcriptional regulation network of Escherichia coli. Nat Genet. 2002;31:64–68. doi: 10.1038/ng881. [DOI] [PubMed] [Google Scholar]
- 90.Segal E., Shapira M., Regev A., Pe'er D., Botstein D., Koller D., et al. Module networks: identifying regulatory modules and their condition-specific regulators from gene expression data. Nat Genet. 2003;34:166–176. doi: 10.1038/ng1165. [DOI] [PubMed] [Google Scholar]
- 91.Huynh-Thu V.A., Irrthum A., Wehenkel L., Geurts P. Inferring regulatory networks from expression data using tree-based methods. PloS One. 2010:5. doi: 10.1371/journal.pone.0012776. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Haury A.C., Mordelet F., Vera-Licona P., Vert J.P. TIGRESS: trustful inference of gene regulation using stability selection. BMC Syst Biol. 2012;6:145. doi: 10.1186/1752-0509-6-145. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Magnusson R., Gustafsson M. LiPLike: towards gene regulatory network predictions of high certainty. Bioinforma (Oxf, Engl) 2020;36:2522–2529. doi: 10.1093/bioinformatics/btz950. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Feizi S., Marbach D., Médard M., Kellis M. Network deconvolution as a general method to distinguish direct dependencies in networks. Nat Biotechnol. 2013;31:726–733. doi: 10.1038/nbt.2635. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Lachmann A., Giorgi F.M., Lopez G., Califano A. ARACNe-AP: gene network reverse engineering through adaptive partitioning inference of mutual information. Bioinforma (Oxf, Engl) 2016;32:2233–2235. doi: 10.1093/bioinformatics/btw216. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Marbach D., Costello J.C., Küffner R., Vega N.M., Prill R.J., Camacho D.M., et al. Wisdom of crowds for robust gene network inference. Nat Methods. 2012;9:796–804. doi: 10.1038/nmeth.2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Malysheva V., Mendoza-Parra M.A., Blum M., Spivakov M., Gronemeyer H. Gene regulatory network reconstruction incorporating 3D chromosomal architecture reveals key transcription factors and DNA elements driving neural lineage commitment. BioRxiv. 2019 doi: 10.1101/303842. [DOI] [Google Scholar]
- 98.Hecker M., Lambeck S., Toepfer S., van Someren E., Guthke R. Gene regulatory network inference: data integration in dynamic models-a review. Biosystems. 2009;96:86–103. doi: 10.1016/j.biosystems.2008.12.004. [DOI] [PubMed] [Google Scholar]
- 99.De Smet R., Marchal K. Advantages and limitations of current network inference methods. Nat Rev Microbiol. 2010;8:717–729. doi: 10.1038/nrmicro2419. [DOI] [PubMed] [Google Scholar]
- 100.Wang Z., Zhang Y., Zang C. BART3D: Inferring transcriptional regulators associated with differential chromatin interactions from Hi-C data. Bioinforma (Oxf, Engl) 2021;37:3075–3078. doi: 10.1093/bioinformatics/btab173. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Ma X., Ezer D., Adryan B., Stevens T.J. Canonical and single-cell Hi-C reveal distinct chromatin interaction sub-networks of mammalian transcription factors. Genome Biol. 2018;19:174. doi: 10.1186/s13059-018-1558-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Gibcus J.H., Samejima K., Goloborodko A., Samejima I., Naumova N., Nuebler J., et al. A pathway for mitotic chromosome formation. Sci (N Y, NY) 2018:359. doi: 10.1126/science.aao6135. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Djekidel M.N., Liang Z., Wang Q., Hu Z., Li G., Chen Y., et al. 3CPET: finding co-factor complexes from ChIA-PET data using a hierarchical Dirichlet process. Genome Biol. 2015;16:288. doi: 10.1186/s13059-015-0851-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Roy S., Siahpirani A.F., Chasman D., Knaack S., Ay F., Stewart R., et al. A predictive modeling approach for cell line-specific long-range regulatory interactions. Nucleic Acids Res. 2015;43:8694–8712. doi: 10.1093/nar/gkv865. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.He B., Chen C., Teng L., Tan K. Global view of enhancer-promoter interactome in human cells. Proc Natl Acad Sci USA. 2014;111:E2191–E2199. doi: 10.1073/pnas.1320308111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Whalen S., Truty R.M., Pollard K.S. Enhancer-promoter interactions are encoded by complex genomic signatures on looping chromatin. Nat Genet. 2016;48:488–496. doi: 10.1038/ng.3539. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Zhang S., Chasman D., Knaack S., Roy S. In silico prediction of high-resolution Hi-C interaction matrices. Nat Commun. 2019;10:5449. doi: 10.1038/s41467-019-13423-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Zhou J., Troyanskaya O.G. Predicting effects of noncoding variants with deep learning-based sequence model. Nat Methods. 2015;12:931–934. doi: 10.1038/nmeth.3547. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Yang B., Liu F., Ren C., Ouyang Z., Xie Z., Bo X., et al. BiRen: predicting enhancers with a deep-learning-based model using the DNA sequence alone. Bioinforma (Oxf, Engl) 2017;33:1930–1936. doi: 10.1093/bioinformatics/btx105. [DOI] [PubMed] [Google Scholar]
- 110.Zhao M., He W., Tang J., Zou Q., Guo F. A hybrid deep learning framework for gene regulatory network inference from single-cell transcriptomic data. Brief Bioinforma. 2022:23. doi: 10.1093/bib/bbab568. [DOI] [PubMed] [Google Scholar]
- 111.Fudenberg G., Kelley D.R., Pollard K.S. Predicting 3D genome folding from DNA sequence with Akita. Nat Methods. 2020;17:1111–1117. doi: 10.1038/s41592-020-0958-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Schwessinger R., Gosden M., Downes D., Brown R.C., Oudelaar A.M., Telenius J., et al. DeepC: predicting 3D genome folding using megabase-scale transfer learning. Nat Methods. 2020;17:1118–1124. doi: 10.1038/s41592-020-0960-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Della Gatta G., Palomero T., Perez-Garcia A., Ambesi-Impiombato A., Bansal M., Carpenter Z.W., et al. Reverse engineering of TLX oncogenic transcriptional networks identifies RUNX1 as tumor suppressor in T-ALL. Nat Med. 2012;18:436–440. doi: 10.1038/nm.2610. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Michoel T., Maere S., Bonnet E., Joshi A., Saeys Y., Van den Bulcke T., et al. Validating module network learning algorithms using simulated data. BMC Bioinforma. 2007;8(Suppl 2):S5. doi: 10.1186/1471-2105-8-s2-s5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Moerman T., Aibar Santos S., Bravo González-Blas C., Simm J., Moreau Y., Aerts J., et al. GRNBoost2 and Arboreto: efficient and scalable inference of gene regulatory networks. Bioinforma (Oxf, Engl) 2019;35:2159–2161. doi: 10.1093/bioinformatics/bty916. [DOI] [PubMed] [Google Scholar]
- 116.Forcato M., Nicoletti C., Pal K., Livi C.M., Ferrari F., Bicciato S. Comparison of computational methods for Hi-C data analysis. Nat Methods. 2017;14:679–685. doi: 10.1038/nmeth.4325. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Hanssen L.L.P., Kassouf M.T., Oudelaar A.M., Biggs D., Preece C., Downes D.J., et al. Tissue-specific CTCF-cohesin-mediated chromatin architecture delimits enhancer interactions and function in vivo. Nat Cell Biol. 2017;19:952–961. doi: 10.1038/ncb3573. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Wang C.Y., Jégu T., Chu H.P., Oh H.J., Lee J.T. SMCHD1 merges chromosome compartments and assists formation of super-structures on the inactive X. Cell. 2018;174(406–21) doi: 10.1016/j.cell.2018.05.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Bintu B., Mateo L.J., Su J.H., Sinnott-Armstrong N.A., Parker M., Kinrot S., et al. Super-resolution chromatin tracing reveals domains and cooperative interactions in single cells. Sci (N Y, NY) 2018:362. doi: 10.1126/science.aau1783. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Sefer E. A comparison of topologically associating domain callers over mammals at high resolution. BMC Bioinforma. 2022;23:127. doi: 10.1186/s12859-022-04674-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Zufferey M., Tavernari D., Oricchio E., Ciriello G. Comparison of computational methods for the identification of topologically associating domains. Genome Biol. 2018;19:217. doi: 10.1186/s13059-018-1596-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Consortium E.P. An integrated encyclopedia of DNA elements in the human genome. Nature. 2012;489:57–74. doi: 10.1038/nature11247. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Ngan C.Y., Wong C.H., Tjong H., Wang W., Goldfeder R.L., Choi C., et al. Chromatin interaction analyses elucidate the roles of PRC2-bound silencers in mouse development. Nat Genet. 2020;52:264–272. doi: 10.1038/s41588-020-0581-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Pang B., Snyder M.P. Systematic identification of silencers in human cells. Nat Genet. 2020;52:254–263. doi: 10.1038/s41588-020-0578-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Cai Y., Zhang Y., Loh Y.P., Tng J.Q., Lim M.C., Cao Z., et al. H3K27me3-rich genomic regions can function as silencers to repress gene expression via chromatin interactions. Nat Commun. 2021;12:719. doi: 10.1038/s41467-021-20940-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Zhu X., Huang L., Li D., Luo J., Huang Q., Wang X., et al. Induce pluripotency via specific distal enhancer-promoter associations. BioRxiv. 2021 doi: 10.1101/2021.06.09.447809. [DOI] [Google Scholar]
- 127.Fudenberg G., Imakaev M., Lu C., Goloborodko A., Abdennur N., Mirny L.A. Formation of chromosomal domains by loop extrusion. Cell Rep. 2016;15:2038–2049. doi: 10.1016/j.celrep.2016.04.085. [DOI] [PMC free article] [PubMed] [Google Scholar]