Skip to main content

3D genomics and its applications in precision medicine

Abstract

Three-dimensional (3D) genomics is an emerging discipline that studies the three-dimensional structure of chromatin and the three-dimensional and functions of genomes. It mainly studies the three-dimensional conformation and functional regulation of intranuclear genomes, such as DNA replication, DNA recombination, genome folding, gene expression regulation, transcription factor regulation mechanism, and the maintenance of three-dimensional conformation of genomes. Self-chromosomal conformation capture (3C) technology has been developed, and 3D genomics and related fields have developed rapidly. In addition, chromatin interaction analysis techniques developed by 3C technologies, such as paired-end tag sequencing (ChIA-PET) and whole-genome chromosome conformation capture (Hi-C), enable scientists to further study the relationship between chromatin conformation and gene regulation in different species. Thus, the spatial conformation of plant, animal, and microbial genomes, transcriptional regulation mechanisms, interaction patterns of chromosomes, and the formation mechanism of spatiotemporal specificity of genomes are revealed. With the help of new experimental technologies, the identification of key genes and signal pathways related to life activities and diseases is sustaining the rapid development of life science, agriculture, and medicine. In this paper, the concept and development of 3D genomics and its application in agricultural science, life science, and medicine are introduced, which provides a theoretical basis for the study of biological life processes.

Introduction

DNA in the nucleus is highly folded and condensed to form chromatin with a specific three-dimensional advanced structure, rather than a single linear structure. This form of existence ensures that DNA remains stable as genetic material. The purpose of genomics is to collectively characterize and quantify all the genes of an organism, and to study their interrelationships and effects on the organism [1]. Genomics mainly studies the structure, function, evolution, localization, and editing of genomes, as well as their effects on organisms. Genomics research has been developing for more than 30 years since the start of the Human Genome Project [2]. In 1990 and 2003, scientists came together to launch the Human Genome Project (HGP) and Encyclopedia of DNA Elements (ENCODE) [3], which respectively define the important genes and linear structure of the human genome and parse all the functional components in the human genome. A series of new and different types of DNA and RNA are revealed and interpreted, including proteins involved in the process of transcription, translation, and other biology, enabling collaboration and information transfer. Therefore, scientists are faced with the new challenge of achieving precise and spatiotemporal individual functions of cells/organisms through different expression regulation modes. Since then, the focus of genomics research has been genome annotation to explore different genes at the genome level as well as the molecular mechanism of labor division and cooperation of each level, namely three-dimensional space structure in different genes, the coding of the interaction between the transcriptional regulatory elements, and the function of genes in a particular cell/biological individual expression regulation of biological effect. These research areas are part of the field of three-dimensional genomics.

Research on the genome has gradually been deepening, starting from the one-dimensional (gene sequence) and two-dimensional (interaction of different sequences) levels and gradually deepening to the three-dimensional (spatial conformation of chromatin) and four-dimensional (variation of sequence over time) levels, which has greatly promoted the development of genomics. Three-dimensional genomics is an emerging discipline that studies precise transcriptional regulation from the whole genome level. It aims to analyze the three-dimensional structure of genomic DNA in organisms, map individual transcriptional interaction networks of organisms, and explore the interaction between regulatory elements and genes, as well as to clarify the specific molecular mechanism of gene expression in time and space to achieve precise capture of gene regulation.

In 2002, with the chromatin conformation capture (3C) [4,5,6] technology coming out, the three-dimensional genome related research entered a new era, after which circular chromatin conformation capture technology (circular chromatin conformation capture technology that can be sequenced) was released after another 4C [5, 6] and chromatin conformation capture carbon copy technology (5C) [5, 7] was applied to multipoint interaction detection. These successively opened the door to the three-dimensional structure of chromatin. In 2015, a global cooperative project-the “4D Nuclear Bodies Project” was launched, and the research of 3D genomics began to enter an era of rapid development. Scientists will be in a few years, or even longer, from space (3D) and time (four-dimensional) perspective study to study the formation principle of cell nucleus structure and explore the influence of cell nucleus on gene expression, cell function, development, and disease occurrence. To have a more in-depth understanding of the popular subject of 3D genomics, this article briefly reviews the research progress of 3D genomics, including the development of 3D genomics technology, the characteristics of 3D genomics technology, and the application of 3D genomics in different fields, to provide strong research evidence and new ideas for the further research of three-dimensional genetics, provide theoretical knowledge for precision biomedicine, and promote the development of human precision medicine.

The birth of 3D genomics and its main technologies

Unlike traditional biology, where the research objects range from “prokaryotes to eukaryotes,” the initial research objects of three-dimensional genomics are mice and humans. The completion of the HGP project in 2002 and the Encyclopedia of DNA Elements Project (ENCODE) [3] project in 2011 allowed scientists to interpret and annotate the characteristics and functions of different sequences in the genome, including coding sequences, noncoding sequences, transcripts, noncoding RNAs, microRNAs, transcription regulatory elements (enhancers, insulators, silencers), transcription factor binding sites, viscous protein binding sites, etc. [3]. At the same time, scientists realized that the spatial structure of the genome is not linearly arranged on the chromosomes. Its 3D structure is important for DNA replication, gene transcription regulation, chromatin condensation and separation, etc. [8]. At the same time (2002), the invention of 3C technology made it possible to use molecular biology methods to explore whether there is an interaction between two target genes, which also laid a solid theoretical foundation for the generation of three-dimensional genomics [9, 10]. In 2009, ChIA-PET [11] and Hi-C [12] technologies were invented. These two technologies allow scientists to elucidate the interactions between different genes and transcriptional regulatory elements at the whole genome level. This has become a milestone in the development of 3D genomics, and it also marks the coming of the era of 3D genomics.

In recent years, innovations in technologies and tools related to three-dimensional genomics research have emerged one after another (Fig. 1). 3C technology-based 4C [4], 5C [7], in situ Hi-C (2013) [13], hybridization probe Capture-Hi-C and CHi-C (2014) [14, 15], DNase Hi-C (2015) [16], DNase Capture-Hi-C [16], and in situ DNase Hi-C (2016) [17] and single-cell Hi-C technology [18] and BL-Hi-C [19] have been invented. These techniques use proximal junctions to detect chromatin interactions between spatially close DNA sequences, which has been applied to analyze the 3D genome structures of various cell types and model organisms, and have found that genomes are generally organized into various hierarchical structural units, including A/B compartments and subcompartments, topological correlation domains (TADs) [20], and chromatin interaction loops [21]. These series of research results show that, in the nucleus, DNA containing one or more genes and regulatory elements can be separated spatially to form topologically associated domains (TAD) [20]. The open or closed areas of chromatin can alternately form functionally related A/B compartments [12, 21]. Due to random collisions, two linearly distant gene loci will recognize the remote regulatory elements at a higher frequency, so as to realize the loop spatial conformation of long-distance transcription regulation, and then form a complex network structure [21, 22]. Therefore, chromatin is a specific spatial conformation formed by orderly folding in a hierarchical form in the nucleus (Fig. 2). [23,24,25]. Chemical crosslinking has assisted in adjacent junction capture technology (CAP-C) [26], DNA adenine methyltransferase (Dam)C [25], etc.

Fig. 1
figure 1

Overview of chromosome conformation capture technology (3C) and 3C-based technology. Formaldehyde fixation is cross-linked, followed by digestion and ligation with specific enzymes. Subsequently, the ligation products are enriched and the contact frequency is determined

Fig. 2
figure 2

The 3D genome’s hierarchical organization. Chromosome territory refers to the region of the nucleus occupied by chromosomes. The entire genome can be partitioned into two multi-MB spatial sections, which can be called “A” and “B” compartments. There are frequent intracompartmental interactions, while there are fewer intracompartmental interactions. “A” compartments: open chromatin that is highly expressed and gene-rich, and has a high GC content, as well as histone indicators for active transcription, and is frequently found in the nucleus’s interior. “B” compartments: closed chromatin with inactive expression, gene deficit, and compact structure, carrying histone markers of gene silence, and located on the nucleus’s periphery. Topological domains are areas of the genome that interact with one another more frequently in chromatin compartments (TAD). These areas are conserved in various cells of different animals and are rich in CTCF and adhesive proteins, and CTCF, cohesive proteins, and mediators mediate the creation of enhancer promoter rings in these locations

Different technologies such as ChIP-Llop, PLAC-seq [22], HiChIP [27], and long-read ChIA-PET [28] based on ChIA-PET [11] technology have also been successfully applied to related research fields. Through remote interactions mediated by specific transcription factors/genes, these studies have discovered a large number of interactions between different transcriptional regulatory elements, enhancers, and genes that determine the identity and status of cells, thereby constructing an intracellular transcriptional regulatory network [29,30,31,32,33] (Table 1). In addition, genome-wide RNA–protein interaction capture technologies have also been developed, such as RIP-seq [34] and CLIP series technologies, including HITS-CLIP, PAR-CLIP, iCLIP, CRAC, etc. [35,36,37]. Recently, the genome-wide DNA–RNA interaction capture technology GRID-seq [38] has also been developed. This technology can provide the most detailed promoter–enhancer-regulated RNA three-dimensional interaction map of the nuclear genome that is closest to the true transcription situation. This technology provides the possibility to explore the DNA–RNA of the whole genome. CRISPER–Cas9-based multicolor fluorescent labeling [39] and in situ biotin-labeled dCas9 technology (CAPTURE) [40] have also been applied in three-dimensional genome research.

Table 1 Characteristics and application of 3D genomics technology

Current innovations in microscopy technology, such as the combined method ChromEMT [41], ensure its position in 3D genomics research. With the successful development of optical microscopes with a resolution of 30 nm, scientists invented a new technology-ChromEMT, by combining chromatin dyes with electron microscopy tomography [42]. These techniques were the first to observe chromatin structure directly in the nucleus through electron microscopy. It was found that chromatin existed in disordered curved chains with a diameter of 5–24 nm at any stage of the cell cycle, but the concentration distribution in three-dimensional space varied with the cell stage. This new discovery “upends” the classical model of chromatin multihierarchy [43]. It is foreseeable that these studies will greatly promote the smooth implementation of the “4D nucleosome project” and will also greatly enhance scientists’ understanding of the relationship between genome-wide spatial conformation and gene transcription regulation [44,45,46]. Indeed, new techniques have been developed to resolve single-cell maps of three-dimensional genomes while preserving information about the spatial location of these cells in their tissue environment, known as the emerging “spatial genomics.” For example, optical reconstruction of chromatin architecture (ORCA) and Hi-M techniques have been applied to Drosophila tissues [47, 48], while multiplexed imaging of nucleome architectures (MINA) is the first method to perform chromatin tracing in mammalian tissues [49], revealing the presence of genomic regions associated with nucleosomal and chromatin markers in different cellular genomes. Moreover, some of these chromatin-associated regions appear to be cell-type specific, while others, primarily nuclear speckle-associated regions, are more conserved across cell types.

Multiplexed error-robust fluorescence in situ hybridization (MERFISH) is applied for high-throughput in situ recognition of thousands of RNA molecules in cells [50], and has been applied to in situ imaging of genomic DNA [51]. Using MERFISH to image TAD structures with ultrahigh-resolution microscopy (STORM) requires further specificity and sensitivity. Secondary hybridization enables MERFISH to trace chromatin structure at super-resolution, which compensates for the defect of the single-cell Hi-C (scHi-C) method’s inability to clearly see the structure of TAD at the single-cell level [52].

To further explore the detailed chromatin structure, researchers have made full use of the fluorescence in situ hybridization (FISH) technique, including exposure to enhancer promoters, but this method is limited to fixed cells and is incapable of elucidating the spatiotemporal genome dynamics. The emergence of genome editing technologies, such as clustered regularly interspaced short palindromic repeats (CRISPR), has revolutionized various techniques in biomedical research, including imaging tools for chromatin interaction to target specific genomic locations of interest in living cells [53, 54]. For example, CRISPR was introduced in living cells by labeling the enhanced green fluorescent protein (EGFP) to the endonuclease-deficient Cas9 (dCas9) protein, which, in combination with imaging techniques, was recruited specifically by small guide RNA (sg) [55]. The method has been applied to other derivatives of the dCas9 linear homolog using different-colored fluorescent protein tags, facilitating the visualization of multiple loci and the distances between them [56, 57]. In addition to dCas9, sgRNA is modified by fusion scaffolds with fluorescently labeled viral RNA recruiting viral proteins, allowing simultaneous multicolor imaging of multiple genomic sites [57,58,59]. These methods are efficient when targeting repetitive sequences in the genome by easily amplifying fluorescent signals, but are not applicable to targeting nonrepetitive genomic regions where signals cannot be strongly amplified for detection. CRISPR-tag and chimeric array of gRNA oligonucleotides (CARGO) have been developed to enhance the signal of dCas9–sgRNA complexes at the nonrepetitive genomic sites [54]. CRISPR-tags are DNA tags that assemble together repeats of two to six CRISPR targets from Cryptococcus that have been knocked into specific human protein-coding genes and recruit dCas9-GFP proteins for imaging [60]. CARGO is capable of delivering 12 guide RNAs cloned in a single plasmid into a single cell, and has been applied to target 2 kb regions of CREs, revealing its dynamics in real-time embryonic stem cell (ESC) [61].

Three-dimensional structure of genome chromatin

Accurate gene expression is a prerequisite for cells and individuals to maintain normal life activities, and the three-dimensional structure of genomic chromatin is the structural basis for precise gene expression regulation. DNA molecules about 2 m in length in mammalian cells are stored in the nucleus with a diameter of about 8 µm in a highly folded and condensed chromatin form, forming a complex and orderly three-dimensional structure, allowing gene expression far away on the linear genome. Regulatory elements can be sufficiently close to their target genes in three-dimensional space, so as to play a fine regulatory role as functional elements. Studies have shown that changes in the three-dimensional structure of genomic chromatin can lead to abnormal gene expression and its regulatory mode, which in turn causes phenotypic changes. Three-dimensional genomics is a research hotspot in the post-genomics era and the post-genome-wide association study (GWAS) era. Compared with the study of two-dimensional linear genome functions, three-dimensional genomics can analyze the precise regulation mechanism of functional elements on gene transcription by exploring the three-dimensional structure of genome chromatin and its regulation of gene expression. Up to now, the three-dimensional structure of chromatin has been resolved in many species, including humans, mice, fruit flies, chickens, parasites, rice, cotton, and corn. In addition, Functional Annotation of Animal Genomes (FAANG) has initially completed the three-dimensional genome study of liver tissues of pigs, chickens, and goats. Due to factors such as DNA sequence, chromosome morphology, number, and tissue specificity of functional elements, the three-dimensional structure of chromatin must be different in the same tissue of different species or in different tissues of the same species. The three-dimensional structure of genomic chromatin is the structural basis for precise gene expression regulation. Bickmore et al. found that the nucleus during the interphase of chromosome division was analyzed and found that there are longer fragments of nuclear chromatin interval and chromosome territory (chromosome territory, CT) in the nucleus, and there are also enhancer–promoter junction regions. These chromatins are three-dimensional. Structure has an important influence on cell gene expression and regulation. With the help of Hi-C and other technologies, researchers have discovered the existence of topologically associated domains (TADs). TADs, as the basic unit of genome folding, exist stably in cells of various species and influence gene expression to a certain degree. The TADs in the nucleus exist in a relatively large structural unit, the chromatin compartment. The chromatin compartment is a manifestation of the apparent state of the genome and is closely related to chromatin activity. Inside TADs, there is a finer folding structure called the chromatin loop (CL), which is usually formed by the interaction of promoters and enhancers (distal) and is the basic function of directly regulating gene expression.

Chromosome territory

Using microscopic observation technology and chromatin conformation capture technology, it is found that each chromosome tends to occupy non-overlapping nuclear regions independently. These regions are called chromosome territories [62, 63]. Within the territory of chromosomes, the genome is not randomly distributed but is related to gene transcription activity. Gene enrichment regions tend to be distributed at the borders of chromosome boundaries, although this is not a common phenomenon. For example, Hoxd gene is activated and expressed in mouse e9.5 embryonic limb buds, but the gene is not transferred to the border of the chromosome when activated [64, 65]. There are also interactions between different chromosome territories, especially at the borders of chromosome territories.

Chromatin compartment A/B

At the Mb level, genomic regions with similar chromosomal characteristics have obvious interactions. For example, there are interactions between the transcriptional activation regions of the genome. These regions usually have higher gene density, chromatin openness, and histone modifications [12]. On the contrary, transcription repression regions, usually gene deserts and heterochromatin regions, will interact with other transcription repression regions. This chromosomal region at the Mb level is called a chromosome compartment, where compartment A is a transcriptional activation area, and compartment B is a transcriptional repression area [12, 66]. This structure was confirmed by Hi-C and microscopic observation techniques [67, 68]. The spatial distribution of chromatin is usually related to various nuclear structures. For example, compartment A is often found in the interior nuclear space, while compartment B is usually located in the nuclear lamina and nucleoli. In human fibroblasts, approximately 40% of the genome is associated with laminin [46].

During the differentiation of mouse embryonic stem cells (mESCs) into neural progenitor cells and further into astrocytes, the interaction pattern between hundreds of gene loci and nuclear lamina gradually changed. The spatial separation of chromatin is not limited to intrachromosomal compartments, but also applies to different interchromosomal compartments [69]. A recent study identified the interaction of compartments between chromosomes. This study found that the interchromosomal transcription-repressive regions are assembled at the nucleolus and usually contain centromeres and ribosomal DNA regions. The finding is consistent with the previously observed nucleolar-related chromosomal domains. On the contrary, the transcription activation region usually assembles at the nuclear speckles. Mechanistically, the location of the B-compartment domain at the periphery of the nucleus depends on the laminin B receptor, laminin A, and laminin C, because the loss of all three proteins will cause the heterochromatin to relocate to the inside of the nucleus [13].

Topological associated domain (TAD)

Another structural level of three-dimensional chromatin is TAD [70, 71]. TAD was originally identified by Hi-C and 5C [72]. In the 2D interaction heat map, it appears as an interaction box on the diagonal. TAD is separated from adjacent regions by a clear boundary, forming an independent regulatory unit, and its main function is to limit the interaction distance of regulatory elements. TAD boundaries are usually highly conservative, but there are also some cell-specific TAD boundaries. TAD borders usually have a large number of chromatin structural proteins CTCF and adhesion proteins (TAD borders in plants generally lack insulating proteins and the borders are not obvious), which play an important role in maintaining the structure and stability of TAD [46, 73, 74]. The TAD boundary not only guides the folding of chromatin into high-level structures, but also correctly guides the long-distance transcriptional regulation. Changes in this boundary can lead to disordered gene regulation. TAD borders usually have histone modifications related to gene activation, such as H3K4me3 and H3K36me3. As the depth of Hi-C sequencing increases [33, 75], using the latest algorithm (Arrowhead algorithm), at a resolution of 5 kb, it is found that the three-dimensional structure of chromatin can be divided into compartmental domains (compartmental domains) and CTCF (CCCTC-binding factor) loops. The formation of the compartment domain has nothing to do with CTCF; it is caused by chromatin state and gene transcription, and the CTCF loop is formed by adhesion protein and CTCF. The compartment area is smaller than TAD, and can be further divided into A/B compartment area [33, 67]. Thus, there may be different compartments in TAD.

Chromatin loop

With the increase of the sequencing depth, more fine interaction peaks are further found in the TAD, which are remotely generated by the regulatory elements in the TAD, which are called interaction loops. Similar to the formation of a loop at the two ends of the TAD, the long-distance interaction between regulatory elements will also make the chromatin loop. Therefore, broadly speaking, the chromatin loop includes TAD and loop structures, and loop is usually hundreds of kilobases, which is much smaller than TAD. Loop is a hotspot in three-dimensional genomics research, which can effectively annotate the interaction of genome functional elements.

Since 2003, the human ENCODE project has revealed hundreds of thousands of genomic functional elements [3]. These regulatory elements play a vital role in the precise expression and regulation of genes. However, the early functional studies of two-dimensional linear genomes could not systematically provide the target gene information of these regulatory elements. Moreover, it is impossible to explain the mechanism through which these regulatory elements interact with target genes that are tens of thousands or even hundreds of thousands of nucleotides away from them. In fact, in the nucleus of eukaryotes, chromatin follows certain rules for complex and orderly three-dimensional folding to form a specific three-dimensional structure of chromatin. This makes the regulatory elements that are far away from the linear genome and their target genes sufficiently close in three-dimensional space, so as to play the fine regulatory role of the functional elements.

With the advancement of the ENCODE project, people realized that this kind of long-distance precise gene expression regulation is widespread in the mammalian genome and is a necessary condition for the normal growth and development of the body. For example, MYC promoter and PVT1 promoter can competitively interact with four enhancers in PVT1 gene. When the PVT1 promoter region is mutated, the interaction between the enhancer and the MYC promoter in the three-dimensional space is enhanced, which promotes the occurrence of cancer [76, 77]. In addition, knocking out the distal enhancer of Sox9 resulted in gender reversal in mice [78].

Chromatin-loop extrusion model

Loop and TAD are both formed by long-distance loops of genomic chromatin, collectively referred to as chromatin loops. Chromatin looping is caused by loop extrusion. The formation of loop is conducive to the interaction of regulatory elements and can regulate gene expression. TAD is usually larger than loop and can limit the interaction of regulatory elements within a certain range. The discovery of TAD and loop is defined with the sequencing depth and the corresponding algorithm, so the results obtained by using different algorithms and resolutions will be different.

According to reports, 86% of chromatin loops are anchored by CTCF [13], and 86% of chromatin loops are anchored by adhesion proteins subunit RAD21 [46]. Adhesin complex can form a ring structure and can move on chromatin. Adhesin can recruit NIPBL and MAU2 proteins and release them from chromatin through WAPL protein. The translocation of adhesion proteins on chromatin requires ATP, because nonspecific inhibition of ATPase or specific mutation of the ATPase domain in the adhesion proteins complex will inhibit this translocation. Gene transcription is also conducive to promoting the translocation of adhesion proteins, which in turn promotes the formation of their circular structures. Mechanistically, CTCF and Adhesin make genomic DNA form TAD through the “chromatin-loop extrusion” model. In this model, adhesion proteins squeeze chromatin outwards until adhesion proteins meet the chromosome boundary formed by CTCF. In this way, several ring structures are formed inside the TAD, which can promote the interaction within the TAD. In the Hi-C interaction matrix, these chromatin loops appear as high-frequency interaction peaks. It is worth noting that the CTCF binding sites at the two boundaries of loop are usually reversed, and their motifs are face to face. Changing the direction of the CTCF motif will destroy the loop and TAD formation. These results strongly indicate that CTCF will promote the formation of loops. In addition, deleting loop extrusion factors, adhesion proteins, or assembly factors NIPBL will cause the widespread disappearance or reduction of TAD and loop. However, there is currently no direct evidence that adhesion proteins can squeeze the chromatin ring. In addition, a recent 4C study found that knocking out CTCF-related TAD boundaries did not affect the mode of local chromatin interaction. Although the study did not evaluate the three-dimensional structure of the whole genome (for example, using 5C or Hi-C for evaluation), in addition to CTCF and adhesion proteins, other factors may also contribute to the formation of TAD [70, 79].

Although TAD is widespread in many species, single-cell Hi-C [80] studies have shown that TAD between single cells is not completely the same. For example, single-cell Hi-C analysis in mice showed that TAD was indeed found in a single cell, but the TAD is different between different cells. After knocking out the adhesion protein, although the TAD boundary of CTCF and the adhesion protein has disappeared, the single cell still has a TAD-like structure [81]. It is not clear whether there is a fundamental difference between TAD and nondependent TAD.

Features of 3D genomics

Spatial structure and three-dimensional transcriptional regulation

Before the emergence of three-dimensional genomics, research methods related to the spatial structure of chromatin or gene interaction were mainly based on low-throughput biological or physical technologies, such as DNA-FISH [9], chromatin karyotype analysis [82], yeast two-hybrid [83], immunoprecipitation [41], protein profile analysis [84], etc. The common features of these technologies are high requirements for pre-preparation work, complicated operations, high background noise, and the inability to more directly obtain the interaction information between different genes [22]. Three-dimensional genomics technology gets rid of these limitations, and enables scientists to directly obtain the DNA sequence and interaction information of different genes and transcriptional regulatory elements with high throughput without prior work. Therefore, scientists can study the molecular mechanism of chromatin spatial conformation, three-dimensional structure of genome, and the regulation of transcription of different genes.

Hi-C technology reveals the interaction between different genes and transcriptional regulatory elements in the whole genome in a “full-to-all” model, and analyzes the interaction mode of all DNA elements in chromatin, which has become a typical representative of three-dimensional genomics technology [12]. The application of this technology has successfully defined chromosome territories and chromatin compartments (including active or inactive regions), and constructed a three-dimensional genome structure with a resolution of 1 Mb [12]. However, short-distance imaging of chromatin interactions remains a technical bottleneck. The key to solving this problem is how to gather more fluorescence into a very short target region. To address this problem, a new technique called Tn5-FISH (Tn5 translocation-based fluorescence in situ hybridization) with a resolution of up to 1 kb has been developed by the research group of Juntao Gao. With the help of Tn5 transposase, it cuts DNA into small fragments with the shortest 40 bp through the “cut and paste” mechanism, while adding splice DNA sequences with known sequences to two segments of the fragment, and constructing probe library by PCR amplification to mark chromatin in the nucleus [85]. On this basis, Bing Ren et al. defined the topological association domains (TADs) and their sizes, and further studied the cooperative relationship between this region and the structural protein factor CTCF (CCCTC-binding factor) [71]. At the same time, they conducted in-depth research on the relatively dynamic sub-TAD, and proposed the concept of chromatin loops, which are considered the smallest unit of gene regulation [5]. Recently, scientists have invented a new 3C method, TiLED-MCC, which comprehensively demonstrates the interaction law and chromatin structure of all regulatory elements in these gene regions of mouse embryonic stem cells [86].

The ChIA-PET technology identifies the functionally related interactions between target genes or transcription factors and different DNA elements in the whole genome, which helps to directly interpret the connections between different interactions of the same gene-specific temporal and spatial expression patterns [12]. As Zhang et al. reported in 2013, using ChIA-PET technology to identify specific antibodies for RNA polymerase II, the study revealed 40,000 long-range interacting enhancer-promoters in three different cell lines [29].

Through the ChIA-PET data mediated by RNA polymerase II and CTCF, the researchers successfully mapped the genome-wide interaction map centered on the promoter, constructed a genome-wide transcriptional regulatory network [70], successfully integrated systematic analysis of different genomics or gene function data with three-dimensional genomics data, and successfully revealed a series of molecular mechanisms related to cancer [11, 87, 88], blood diseases [89, 90], noncoding RNA function [91], sex determination [78], and body immunity [92, 93].

There is no doubt that these latest research findings have laid a solid foundation for the further study of the three-dimensional transcriptional regulatory network of eukaryotic genomes. In the future, three-dimensional genomics research will be more concentrated in the construction of gene three-dimensional network, the discovery of key transcription factors/regulatory elements, and the biological effects of gene interaction on cells/individuals.

Integration of multiple omics data

The main content of three-dimensional genomics research is the spatial conformation of the genome. By studying the spatial structure relationship between different genes and transcriptional regulatory elements [94], the cooperative relationship between different genes and transcriptional regulatory elements can be revealed. In short, its main content is to reveal the relationship between different elements in the genome.

To accurately analyze the influence or biological effects of the relationship between different elements in the genome on the functions of different genes, it is necessary to make detailed annotations on the biological effects of different genes, transcriptional regulatory elements, or other biological macromolecules involved in the transcriptional regulation of specific genes. Through the biological process of dynamic time and space, the biological effects of the spatial structure, including the downstream RNA transcription and expression pattern and the level of protein translation regulation, are used for in-depth research. In the process of studying the spatial structure of the three-dimensional genome and its influence on the transcriptional regulation of different genes, it is necessary to integrate other different types of omics data to define and annotate different elements in the whole genome. Only by combining the biological processes of temporal and spatial dynamics can scientists accurately and deeply carry out the next step in the interpretation of the relationship between gene interaction and function [95]. Zhao et al. analyzed pig tissues with a combination of different techniques, including ChIP-seq, Hi-C, RNA-seq, and ATAC-seq research technical systems. In this study, 199 groups of data containing 12 kinds of epigenetic regulation were obtained from Meishan pig, Duroc pig, lean big white pig and lard pig, and the genome of specific regulatory elements in pig tissue was analyzed. These findings serve as a foundation for future research on functional genomics and trait expression regulatory mechanisms.

At present, data related to gene expression and gene annotation in the genome are often used in this type of analysis, such as RNA-seq data that reveal gene expression levels in the genome, and ChIP-seq data that identify different nucleosomal properties or transcription factor binding sites, and GWAS data indicating possible biological significance in the whole genome [96,97,98]. The integration of these studies with 3D genomics data has deepened our understanding of existing gene functions and modes of action. However, integrating multiple data is an enormous challenge. Ideally, the integration of relevant studies requires multi-omics data covering all biological levels, intermediate steps from DNA to protein, all developmental stages from stem cells to death, and all cell types from stem cells to functional cells. In fact, only a small fraction of the required data is generated, and more data will be available in the future. Spatial transcriptome, Hi-C data will contribute to a better understanding of gene transcription and its regulation. In addition, histological data are sensitive to sex, ethnic genetic background, and other variation, the effects of which are not well presented in most public databases. Data on ethnic diversity remain generally unavailable. In 2019, researchers found that about 78% of GWAS individuals were of European ancestry [99]. The diversity of brain histology data is even smaller. For example, in the current version of the Accelerating Medicines Partnership Parkinson's Disease (AMP-PD), less than 4% of participants were non-white/Caucasian [100]. This is a huge problem for research aimed at being more inclusive. Projects such as the African Descent Neuroscience Study promise to fill this gap. The integration of data relies on large samples provided by research laboratories, but the number of samples is still far less than the number of traits. Before integrating multi-omics, researchers must reduce the dimensionality. Related studies have shown that classification based on weighted gene co-expression network analysis (WGCNA) co-expression modules can better cope with differences between datasets than classification based on raw gene expression. Other dimensionality reduction techniques such as support vector machines (SVM), random forests (RF), and singular value decomposition (SVD) are also commonly used to reduce overfitting problems. Advanced deep learning methods such as variational autoencoders (VAE) also output low-dimensional potential representations of high-dimensional data. Multiple test information and saliency criteria are concomitant issues.

Applications of 3D genomics

Interpreting the function of genes

Before the emergence of three-dimensional genomics, the main methods of gene function research were gene knockout, gene knockdown, gene fusion, and association analysis of population traits. The progress of related work mainly depends on whether the changes in gene phenotype or biological traits can be observed. If no changes in the relevant traits are observed, the research work will be suspended or suspended. The objects of these research work focus on the gene itself, and the research on the upstream and downstream pathways of gene function is mainly based on the support of different experimental hypotheses or indirect evidence. For example, Chen et al. reported in 2008 that, by integrating the binding sites of 13 different transcription factors in the embryonic stem cell genome, they found a series of binding sites related to gene transcription regulation activity. This work provides a new way for gene function and genome annotation, and has received extensive attention from scientists. However, due to the lack of the interaction between these binding sites and the target gene, and the limitations of technical conditions, the results of this study have not been widely used in the study of gene function [101].

In three-dimensional genomics, the identification of the mode of interaction between transcription factors and regulatory elements will clear the way for similar research. For example, in human red blood cells, researchers have found that, in the absence of the GATA1 gene, the β-globulin promoter region can interact with locus control regions (LCR) to enhance gene expression [102]. Further studies have shown that, through the interaction between LCR and γ-globulin, the expression level of γ-globulin in the fetal period can be increased to 85% of the total globulin level [103]. The above results indicate that gene interactions dominated by three-dimensional spatial structure can determine developmental pathways or fate [104]. By determining the interconnection between these binding sites and target genes, scientists can further study the on/off of transcriptional regulation mode and the high/low level of gene expression under dynamic spatiotemporal conditions.

Screening of key transcription factors or genes

The ultimate goal of genome-wide spatial conformation research is to study gene functions in depth and provide guidance or research clues for subsequent experiments or analysis. As a carrier of cell/biological individual genetic function information, the study of gene function or transcriptional regulation mode has always been the focus of scientific researchers. Especially the research on the functions and dynamic transcription regulation modes of transcription factors/key genes that have a significant impact on the biological characteristics of cells/biological individuals.

After obtaining the gene interaction/cooperation information on the three-dimensional structure, the relevant data can be used to identify, classify, and evaluate the biological effects of different types of gene interaction/cooperation data. These works will help researchers to further study gene interaction/assistance. At the same time, these research results can be integrated into the three-dimensional structure of the whole genome again.

Researchers can use these data to better analyze and identify the biological effects of different gene interactions/cooperation in the whole genome, and it is also helpful to identify different transcription factors/key genes at specific developmental stages, so as to further explore gene functions. For example, Zhang et al. reported in 2013 that, through improved ChIA-PET technology, three-dimensional transcriptional regulatory networks were constructed in embryonic stem cells and neural stem cells, and these networks confirmed the hypothesis that organisms/cells may have transcription factories. At the same time, some transcription factors/key genes such as Oct4, Sox2, Klf4, and c-Myc were found to exist in this network, and their interacting genes and transcriptional regulatory elements have also been successfully verified [29].

In other different research fields, certain transcription factors/key genes have also attracted the attention of many scientists, but their genome-wide gene interaction/assistance relationship is still unclear, for example, MyoD and Myf5 in the differentiation and development of muscle cells [105], Sox2 and Sox gene families in the development of nerve cells [106], IFN and IL gene families in immune cells [107], Ras and p53 in tumor cells [108], and Hox gene families in the development process [109]. How do these genes interact with other transcription factors/genes/regulatory elements? Do they also have a genome-wide transcriptional regulatory network like mouse embryonic stem cells? What role do they play in the dynamic development of cells/biological entities? These issues are awaiting continued attention and in-depth research by scientists in their future work.

Enhancers and super enhancers

Enhancers are noncoding DNA elements that regulate gene function and transcription regulation through long-distance interaction, and are also one of the most well-known research objects in the field of three-dimensional genomics. They are mainly located in the open chromatin region (DNase I-sensitive part or ATAC-seq signal region), which can increase gene expression. In addition, the DNA sequence of the enhancer is rich in the binding elements of cell-specific transcription factors, and is rich in epigenetic regulatory factors such as transcription activator CBP/P300 and histone-modifying enzymes [110]. Enhancers are widely present in the whole genome and have strong spatiotemporal specificity. They can regulate the transcriptional regulation of target genes through long-distance interaction with target promoters [111]. Scientists have been interested in the study of enhancers for a long time, but due to the limitations of technology and scientific research, the study of enhancers has been stuck in the isolation and identification stage, lacking relevant theories and systematic research support. The development of three-dimensional genomics has injected a strong momentum into enhancer research. For example, Zhang et al.’s study revealed 40,000 long-range enhancer–promoter interactions in three different cell lines. Further analysis also showed that 40% of enhancers interacted with their neighboring genes, and the target genes of the remaining 60% enhancers were clearly identified [29].

In 2013, the Richard Young team proposed the concept of super enhancer [112]. On the basis of systematic analysis of enhancer data of different cells/developmental stages, they identified a class of important organisms in the process of life, including cell identity, cell fate, and growth cycle. The cluster of enhancers related to the learning process is named super-enhancer. Such elements can be identified in any cell, can define cell identity, are very sensitive to environmental changes, and have an important impact on the expression of proto-oncogenes [113, 114]. Some important traits and disease-related mutations are mostly located in the super-enhancer region, which can form the super-enhancer domain and express enhancer RNA [115, 116]. At present, there are many controversies about the concept of super-enhancers [117], but because it satisfies people’s needs for narrowing the research candidates, the research on super-enhancers has made explosive progress [118, 119]. For example, in cancer research, super-enhancers are used as a new type of drug target [120]; in cell research, they are used as candidates for nuclear microenvironment research [121]; in immune cells, they are used as key factors in developmental regulation [122, 123]. However, the specific composition, regulatory mechanism, and biological functions of super enhancers need to be further analyzed. There is no doubt that three-dimensional genomics technology is a powerful tool to analyze this series of problems, and its progress will lay the foundation for the study of accurately revealing the regulatory mechanism of super-enhancers.

Precision biology and 3D genomics

The precision of precision biology is mainly reflected in the two aspects of resource conservation/maximization of effects and accurate targeting, and this must be based on an in-depth understanding of the underlying molecular mechanisms of existing biological processes/traits. The source of the concept of precision biology is quoted from precision medicine, and its foundation is based on the theory of big data, through the integration of genome, proteomics, and other omics technologies and biological processes, as well as the analysis, identification, verification, and application of biomarkers for large sample groups and specific trait types. The next step is to accurately find the target genes that determine the traits, and formulate personalized improvement strategies to accurately classify the dynamic spatiotemporal expression patterns of specific traits, and finally achieve the goal of improving and enhancing traits/biological individuals. In recent years, the scientific community has launched a series of precision biology projects. However, due to the lack of knowledge and understanding of basic gene functions and molecular mechanisms of action, the current progress tends to be slow [124, 125]. People have realized that big data should not only be used as the primary factor of precision biology; more attention should be paid to functional information, and the role of biological macromolecules in precision biology should be strengthened [126].

Three-dimensional genomics can reveal the general law of transcriptional regulation of a certain gene in the whole genome and, on this basis, further clarify the regulatory elements required in the transcription process of a specific gene, as well as the mutual cooperation relationship with other different genes, which is a single study of gene function, and molecular mechanism of action provides a brand-new approach [29]. Studies have shown that, when genes interact with different transcription elements/protein molecules, they can present different transcription patterns such as expression/closure. Further studies have shown that, when the concentration of transcription elements/protein molecules in the gene expression pattern changes, the level of gene expression also continuously increases or decreases [127]. The results of this study potentially indicate that the interactions between different genes/transcriptional elements in organisms/cells revealed by three-dimensional genomics can be used to determine the spatiotemporal expression patterns of genes under specific spatiotemporal conditions. The difference in their interaction mode and expression level will directly affect the level of expression of the gene or its temporal and spatial specificity. Therefore, the small changes in the expression pattern of the gene in the nucleus caused by the interaction of these genes/transcriptional elements can be directly used in the research of precision biology, and will directly promote the development of precision biology.

Conclusions and prospects

In summary, through three-dimensional genomics, we can clearly understand the mechanism of spatial conformation of chromatin, the mechanism of gene transcription regulation, the formation mechanism of biological traits, the transmission mechanism of signal pathways, and the operating mechanism of the genome. The combination of 3C technology and other derivative sequencing technologies applied to three-dimensional genomics research has increasingly clarified the spatial conformation of the three-dimensional genomes of animals, plants, and microorganisms. Although related progress has been amazing in recent years, research on 4D nucleosomes is still at the tip of the iceberg [127]. Large improvements are still needed in experimental methods and calculation tools, and relevant technical standards have not yet been formally established. This has led to uneven data quality, and related hypotheses and theories still need to be improved. Because of the lack of common standards for the operation of testing experiments, the results of research cannot be directly compared. For example, the 5C technique allows a more comprehensive analysis of chromatin-loop interactions for a numerous genes by mapping interactions between multiple motifs in parallel, and is superior to 3C in statistical analysis, but is limited by the models and assumptions used to define the expected frequency of interactions [7]. Studies based on tomographic electron microscopy and combined labeling technology (ChromEMT) show that the three-dimensional structure of the nucleus shows that the accumulation of chromatin in the nucleus is based on the concentration of chromatin, which directly denies the chromatin classification proposed by the Hi-C technology structural hypothesis. It is believed that this aggregation of different concentrations determines the activity and recognition mechanism of DNA [128]. Nevertheless, the development of single-cell, high-throughput, and multi-omics 3D genomic analyses will allow us to understand in as much detail as possible the dynamic structure of the genomic organization under normal and diseased conditions. It will facilitate further exploration of gene expression regulation patterns, transcriptional–translational processes, and synergistic patterns of different genes.

There is a proliferation of emerging technologies, and the integration of emerging technologies with the 3D genome is becoming closer. Researchers developed CAPTURE technology based on proteomic and 3C interactions, as well as multi-ChIA technology based on single-molecule and single-cell principles [129]. At the same time, the ability of researchers to integrate different types of data (such as chromatin interaction data and imaging-based distance measurement) is still very limited, and there is a lack of methods that can measure and explain the differences in chromosomal and nuclear structure of different cells. The development of 3D genomics requires the establishment of a high-throughput technology-based gene function validation platform to verify data and different hypotheses, so as to establish research standards and technical platforms for 3D genomics.

The establishment of platforms and standards will help scientists to accumulate high-quality data more effectively and improve relevant theories, so that the development of 3D genomics will enter a virtuous cycle, while the development of 3D genomics will also provide a strong linkage and bridge for interdisciplinary research.

In addition, because of the large scale of data involved in the current research process of three-dimensional genomics, a series of different types of omics data are involved in the integration analysis. Commonly used visualization tools such as Genome Browser and DNA Database cannot well display the inherent characteristics and associations of 3D genomics data [130]. Therefore, a series of new visualization tools have been developed and used. These tools have their own emphasis and characteristics when displaying 3D genomics data, such as juicebox [131], 3D-genome Interaction Viewer and database [132], epigenome browser [133], and 3Disease browser [94]. Three-dimensional genome data show the interrelationships of different genome spatial conformations and different gene functions in multiple dimensions. However, the current tools cannot clarify the relationship among the three levels of interaction strength, spatial conformation, and gene synergy in 3D genome data. Especially at the user level, there are still major deficiencies in the usability and experience of terminal researchers; the newly developed three-dimensional genomics tools still have much room for improvement in terms of user experience and usability. The above shortcomings depend on further improvement in future work.

Finally, the development of precision biology depends on scientists’ perfection of a series of basic theories such as gene function, expression regulation mode, transcription and translation process, and synergy mode of different genes at the whole genome level. There is also a lack of a clear understanding of the molecular mechanism of the relationship between chromatin conformation and biological processes in the nucleus (including transcription, DNA replication, and chromosome separation). However, the development of three-dimensional genomics can provide a powerful bridge for the dynamic research of different biological processes. The gap between these basic biological phenomena and people’s cognition can be filled by means of high-level collaboration and multidisciplinary integration. This requires collaboration between research groups that specialize in imaging, genomics, computer science, and physics. However, the research of 3D genomics is still in its infancy, and how to transform the current obtained phased data into research results with clinical application value, and truly promote the development of related fields, is the top priority of 3D genomics research, such as gradually understanding the process of cancer and other diseases as well as chromatin internal and interchromatin structural changes. According to these changes, therapeutic targets can be found and targeted drugs developed, so as to regulate chromatin interactions for treating diseases. After the gene regulation network and mechanism of complex traits of animals and plants has been elucidated, how to use these advantages of three-dimensional structural characteristics to improve the level of biotechnology and improve the quality of life of people is a scientific question that needs to be explored by scientists. With the rapid development of single-cell, high-throughput, and multi-omics 3D genomic analysis, researchers are able to obtain deeper biological information to support the development of human society and the improvement of healthcare.

Availability of data and materials

Not applicable.

Abbreviations

3D:

Three-dimensional

HGP:

Human Genome Project

ENCODE:

Encyclopedia of DNA elements

3C:

Chromatin conformation capture

4C:

Circular chromatin conformation capture

5C:

Chromatin conformation capture carbon copy

Hi-C:

High-throughput chromosome conformation capture

ChromEMT:

ChromEM tomography

GWAS:

Genome-wide association study

FAANG:

Functional Annotation of Animal Genomes

CT:

Chromosome territory

CL:

Chromatin loop

TADs:

Topologically associated domains

mESC:

Mouse embryonic stem cells

FISH:

Fluorescence in situ hybridization

CHIP:

Chromatin immunoprecipitation

LCR:

Locus control regions

SPRITE:

Split-pool recognition of interactions by tag extension

GAM:

Genome architecture mapping

CAP-C:

Chemical-crosslinking assisted proximity capture

Dam:

DNA adenine methyltransferase

MINA:

Multiplexed imaging of nucleome architectures

ORCA:

Optical reconstruction of chromatin architecture

scHi-C:

Single-cell Hi-C

CRISPR:

Clustered regularly interspaced short palindromic repeats

EGFP:

Enhanced green fluorescent protein

WGCNA:

Weighted gene co-expression network analysis

References

  1. Research WHOACoH. Genomics and world health: report of the Advisory Committee on Health Research. Publications of World Health Organization 2002.

  2. Green ED, Watson JD, Collins FS. Human Genome Project: Twenty-five years of big biology. Nature. 2015;526:29–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. The ENCODE (ENCyclopedia Of DNA Elements) Project. Science (New York, NY). 2004; 306:636–640.

  4. Splinter E, de Laat W. The complex transcription regulatory landscape of our genome: control in three dimensions. EMBO J. 2011;30:4345–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Yu M, Ren B. The three-dimensional organization of mammalian genomes. Annu Rev Cell Dev Biol. 2017;33:265–89.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Birney E. The international Human Genome Project. Hum Mol Genet. 2021;30:R161–3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Dostie J, Richmond TA, Arnaout RA, Selzer RR, Lee WL, Honan TA, Rubio ED, Krumm A, Lamb J, Nusbaum C, et al. Chromosome Conformation Capture Carbon Copy (5C): a massively parallel solution for mapping interactions between genomic elements. Genome Res. 2006;16:1299–309.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Sexton T, Cavalli G. The 3D genome shapes up for pluripotency. Cell Stem Cell. 2013;13:3–4.

    Article  CAS  PubMed  Google Scholar 

  9. Langer-Safer PR, Levine M, Ward DC. Immunological method for mapping genes on Drosophila polytene chromosomes. Proc Natl Acad Sci USA. 1982;79:4381–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. de Wit E, de Laat W. A decade of 3C technologies: insights into nuclear organization. Genes Dev. 2012;26:11–24.

    Article  PubMed  PubMed Central  Google Scholar 

  11. Fullwood MJ, Liu MH, Pan YF, Liu J, Xu H, Mohamed YB, Orlov YL, Velkov S, Ho A, Mei PH, et al. An oestrogen-receptor-alpha-bound human chromatin interactome. Nature. 2009;462:58–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Lieberman-Aiden E, van Berkum NL, Williams L, Imakaev M, Ragoczy T, Telling A, Amit I, Lajoie BR, Sabo PJ, Dorschner MO, et al. Comprehensive mapping of long-range interactions reveals folding principles of the human genome. Science (New York, NY). 2009;326:289–93.

    Article  CAS  Google Scholar 

  13. Rao SSP, Huntley MH, Durand NC, Stamenova EK, Bochkov ID, Robinson JT, Sanborn AL, Machol I, Omer AD, Lander ES, Aiden EL. A 3D map of the human genome at kilobase resolution reveals principles of chromatin looping. Cell. 2014;159:1665–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Schoenfelder S, Sexton T, Chakalova L, Cope NF, Horton A, Andrews S, Kurukuti S, Mitchell JA, Umlauf D, Dimitrova DS, et al. Preferential associations between co-regulated genes reveal a transcriptional interactome in erythroid cells. Nat Genet. 2010;42:53–61.

    Article  CAS  PubMed  Google Scholar 

  15. Hughes JR, Roberts N, McGowan S, Hay D, Giannoulatou E, Lynch M, De Gobbi M, Taylor S, Gibbons R, Higgs DR. Analysis of hundreds of cis-regulatory landscapes at high resolution in a single, high-throughput experiment. Nat Genet. 2014;46:205–12.

    Article  CAS  PubMed  Google Scholar 

  16. Ma W, Ay F, Lee C, Gulsoy G, Deng X, Cook S, Hesson J, Cavanaugh C, Ware CB, Krumm A, et al. Fine-scale chromatin interaction maps reveal the cis-regulatory landscape of human lincRNA genes. Nat Methods. 2015;12:71–8.

    Article  PubMed  Google Scholar 

  17. Ramani V, Cusanovich DA, Hause RJ, Ma W, Qiu R, Deng X, Blau CA, Disteche CM, Noble WS, Shendure J, Duan Z. Mapping 3D genome architecture through in situ DNase Hi-C. Nat Protoc. 2016;11:2104–21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Stevens TJ, Lando D, Basu S, Atkinson LP, Cao Y, Lee SF, Leeb M, Wohlfahrt KJ, Boucher W, O’Shaughnessy-Kirwan A, et al. 3D structures of individual mammalian genomes studied by single-cell Hi-C. Nature. 2017;544:59–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Liang Z, Li G, Wang Z, Djekidel MN, Li Y, Qian M-P, Zhang MQ, Chen Y. BL-Hi-C is an efficient and sensitive approach for capturing structural and regulatory chromatin interactions. Nat Commun. 2017;8:1622.

    Article  PubMed  PubMed Central  Google Scholar 

  20. Risca VI, Greenleaf WJ. Unraveling the 3D genome: genomics tools for multiscale exploration. Trends Genet. 2015;31:357–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Schmitt AD, Hu M, Ren B. Genome-wide mapping and analysis of chromosome architecture. Nat Rev Mol Cell Biol. 2016;17:743–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Risca VI, Greenleaf WJ. Unraveling the 3D genome: genomics tools for multiscale exploration. Trends Genet TIG. 2015;31:357–72.

    Article  CAS  PubMed  Google Scholar 

  23. Quinodoz SA, Bhat P, Chovanec P, Jachowicz JW, Ollikainen N, Detmar E, Soehalim E, Guttman M. SPRITE: a genome-wide method for mapping higher-order 3D interactions in the nucleus using combinatorial split-and-pool barcoding. Nat Protoc. 2022;17:36–75.

    Article  CAS  PubMed  Google Scholar 

  24. Zheng M, Tian SZ, Capurso D, Kim M, Maurya R, Lee B, Piecuch E, Gong L, Zhu JJ, Li Z, et al. Multiplex chromatin interactions with single-molecule precision. Nature. 2019;566:558–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Redolfi J, Zhan Y, Valdes-Quezada C, Kryzhanovska M, Guerreiro I, Iesmantavicius V, Pollex T, Grand RS, Mulugeta E, Kind J, et al. DamC reveals principles of chromatin folding in vivo without crosslinking and ligation. Nat Struct Mol Biol. 2019;26:471–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. You Q, Cheng AY, Gu X, Harada BT, Yu M, Wu T, Ren B, Ouyang Z, He C. Direct DNA crosslinking with CAP-C uncovers transcription-dependent chromatin organization at high resolution. Nat Biotechnol. 2021;39:225–35.

    Article  CAS  PubMed  Google Scholar 

  27. Mumbach MR, Rubin AJ, Flynn RA, Dai C, Khavari PA, Greenleaf WJ, Chang HY. HiChIP: efficient and sensitive analysis of protein-directed genome architecture. Nat Methods. 2016;13:919–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Li X, Luo OJ, Wang P, Zheng M, Wang D, Piecuch E, Zhu JJ, Tian SZ, Tang Z, Li G, Ruan Y. Long-read ChIA-PET for base-pair-resolution mapping of haplotype-specific chromatin interactions. Nat Protoc. 2017;12:899–915.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Zhang Y, Wong C-H, Birnbaum RY, Li G, Favaro R, Ngan CY, Lim J, Tai E, Poh HM, Wong E, et al. Chromatin connectivity maps reveal dynamic promoter–enhancer long-range associations. Nature. 2013;504:306–10.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Grubert F, Srivas R, Spacek DV, Kasowski M, Ruiz-Velasco M, Sinnott-Armstrong N, Greenside P, Narasimha A, Liu Q, Geller B, et al. Landscape of cohesin-mediated chromatin loops in the human genome. Nature. 2020;583:737–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Tang L, Hill MC, Wang J, Wang J, Martin JF, Li M. Predicting unrecognized enhancer-mediated genome topology by an ensemble machine learning model. Genome Res. 2020;30:1835–45.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Bhattacharyya S, Chandra V, Vijayanand P, Ay F. Identification of significant chromatin contacts from HiChIP data by FitHiChIP. Nat Commun. 2019;10:4221.

    Article  PubMed  PubMed Central  Google Scholar 

  33. Zhao Y, Hou Y, Xu Y, Luan Y, Zhou H, Qi X, Hu M, Wang D, Wang Z, Fu Y, et al. A compendium and comparative epigenomics analysis of cis-regulatory elements in the pig genome. Nat Commun. 2021;12:2217.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Keene JD, Komisarow JM, Friedersdorf MB. RIP-ChIP: the isolation and identification of mRNAs, microRNAs and protein components of ribonucleoprotein complexes from cell extracts. Nat Protoc. 2006;1:302–7.

    Article  CAS  PubMed  Google Scholar 

  35. Bartel T, Khalil M, Traina M, Mihaljevic T, Suri RM. Percutaneous mitral annuloplasty complements clip implantation in functional mitral regurgitation. Eur Heart J. 2019;40:2584.

    Article  PubMed  Google Scholar 

  36. Zhang Y, Xie S, Xu H, Qu L. CLIP: viewing the RNA world from an RNA–protein interactome perspective. Sci China Life Sci. 2015;58:75–88.

    Article  CAS  PubMed  Google Scholar 

  37. Lee FCY, Ule J. Advances in CLIP technologies for studies of protein–RNA interactions. Mol Cell. 2018;69:354–69.

    Article  CAS  PubMed  Google Scholar 

  38. Özdemir V, Patrinos GP. David Bowie and the art of slow innovation: a fast-second winner strategy for biotechnology and precision medicine global development. OMICS. 2017;21:633–7.

    Article  PubMed  Google Scholar 

  39. Gupta D, Bhattacharjee O, Mandal D, Sen MK, Dey D, Dasgupta A, Kazi TA, Gupta R, Sinharoy S, Acharya K, et al. CRISPR-Cas9 system: a new-fangled dawn in gene editing. Life Sci. 2019;232: 116636.

    Article  CAS  PubMed  Google Scholar 

  40. Kim H, Lee S, Seo HW, Kang B, Moon J, Lee KG, Yong D, Kang H, Jung J, Lim E-K, et al. Clustered regularly interspaced short palindromic repeats-mediated surface-enhanced Raman scattering assay for multidrug-resistant bacteria. ACS Nano. 2020;14:17241.

    Article  CAS  PubMed  Google Scholar 

  41. Niepel M, Strambio-de-Castillia C, Fasolo J, Chait BT, Rout MP. The nuclear pore complex-associated protein, Mlp2p, binds to the yeast spindle pole body and promotes its efficient assembly. J Cell Biol. 2005;170:225–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Giles AJ, Dai S, Vurgaftman I, Hoffman T, Liu S, Lindsay L, Ellis CT, Assefa N, Chatzakis I, Reinecke TL, et al. Ultralow-loss polaritons in isotopically pure boron nitride. Nat Mater. 2018;17:134–9.

    Article  CAS  PubMed  Google Scholar 

  43. Ou HD, Phan S, Deerinck TJ, Thor A, Ellisman MH, O'Shea CC. ChromEMT: Visualizing 3D chromatin structure and compaction in interphase and mitotic cells. Science. 2017; 357.

  44. Li G. The 3D organization of genome in the nucleus: from the nucleosome to the 4D nucleome. Sci China Life Sci. 2020;63:791–4.

    Article  PubMed  Google Scholar 

  45. Sparks TM, Harabula I, Pombo A. Evolving methodologies and concepts in 4D nucleome research. Curr Opin Cell Biol. 2020;64:105–11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Rao SSP, Huang S-C, Glenn St Hilaire B, Engreitz JM, Perez EM, Kieffer-Kwon K-R, Sanborn AL, Johnstone SE, Bascom GD, Bochkov ID, et al. Cohesin loss eliminates all loop domains. Cell. 2017;171:305.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Cardozo Gizzi AM, Cattoni DI, Fiche JB, Espinola SM, Gurgo J, Messina O, Houbron C, Ogiyama Y, Papadopoulos GL, Cavalli G, et al. Microscopy-based chromosome conformation capture enables simultaneous visualization of genome organization and transcription in intact organisms. Mol Cell. 2019;74:212-222.e215.

    Article  CAS  PubMed  Google Scholar 

  48. Mateo LJ, Murphy SE, Hafner A, Cinquini IS, Walker CA, Boettiger AN. Visualizing DNA folding and RNA in embryos at single-cell resolution. Nature. 2019;568:49–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Liu M, Lu Y, Yang B, Chen Y, Radda JSD, Hu M, Katz SG, Wang S. Multiplexed imaging of nucleome architectures in single cells of mammalian tissue. Nat Commun. 2020;11:2907.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Chen KH, Boettiger AN, Moffitt JR, Wang S, Zhuang X. RNA imaging spatially resolved, highly multiplexed RNA profiling in single cells. Science. 2015;348:aaa6090.

    Article  PubMed  PubMed Central  Google Scholar 

  51. Su JH, Zheng P, Kinrot SS, Bintu B, Zhuang X. Genome-scale imaging of the 3D organization and transcriptional activity of chromatin. Cell. 2020;182:1641-1659.e1626.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Kim HJ, Yardımcı GG, Bonora G, Ramani V, Liu J, Qiu R, Lee C, Hesson J, Ware CB, Shendure J, et al. Capturing cell type-specific chromatin compartment patterns by applying topic modeling to single-cell Hi-C data. PLoS Comput Biol. 2020;16: e1008173.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Kempfer R, Pombo A. Methods for mapping 3D chromosome architecture. Nat Rev Genet. 2020;21:207–26.

    Article  CAS  PubMed  Google Scholar 

  54. Lakadamyali M, Cosma MP. Visualizing the genome in high resolution challenges our textbook understanding. Nat Methods. 2020;17:371–9.

    Article  CAS  PubMed  Google Scholar 

  55. Chen B, Gilbert LA, Cimini BA, Schnitzbauer J, Zhang W, Li GW, Park J, Blackburn EH, Weissman JS, Qi LS, Huang B. Dynamic imaging of genomic loci in living human cells by an optimized CRISPR/Cas system. Cell. 2013;155:1479–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Ma H, Naseri A, Reyes-Gutierrez P, Wolfe SA, Zhang S, Pederson T. Multicolor CRISPR labeling of chromosomal loci in human cells. Proc Natl Acad Sci U S A. 2015;112:3002–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Shao S, Zhang W, Hu H, Xue B, Qin J, Sun C, Sun Y, Wei W, Sun Y. Long-term dual-color tracking of genomic loci by modified sgRNAs of the CRISPR/Cas9 system. Nucleic Acids Res. 2016;44: e86.

    Article  PubMed  PubMed Central  Google Scholar 

  58. Fu Y, Rocha PP, Luo VM, Raviram R, Deng Y, Mazzoni EO, Skok JA. CRISPR-dCas9 and sgRNA scaffolds enable dual-colour live imaging of satellite sequences and repeat-enriched individual loci. Nat Commun. 2016;7:11707.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Ma H, Tu LC, Naseri A, Huisman M, Zhang S, Grunwald D, Pederson T. Multiplexed labeling of genomic loci with dCas9 and engineered sgRNAs using CRISPRainbow. Nat Biotechnol. 2016;34:528–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Chen B, Zou W, Xu H, Liang Y, Huang B. Efficient labeling and imaging of protein-coding genes in living cells using CRISPR-Tag. Nat Commun. 2018;9:5065.

    Article  PubMed  PubMed Central  Google Scholar 

  61. Gu B, Swigut T, Spencley A, Bauer MR, Chung M, Meyer T, Wysocka J. Transcription-coupled changes in nuclear mobility of mammalian cis-regulatory elements. Science. 2018;359:1050–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Cardozo Gizzi AM, Cattoni DI, Fiche J-B, Espinola SM, Gurgo J, Messina O, Houbron C, Ogiyama Y, Papadopoulos GL, Cavalli G, et al. Microscopy-based chromosome conformation capture enables simultaneous visualization of genome organization and transcription in intact organisms. Mol Cell. 2019; 74.

  63. Misteli T. The self-organizing genome: principles of genome architecture and function. Cell. 2020;183:28–45.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Bianco S, Annunziatella C, Andrey G, Chiariello AM, Esposito A, Fiorillo L, Prisco A, Conte M, Campanile R, Nicodemi M. Modeling single-molecule conformations of the HoxD region in mouse embryonic stem and cortical neuronal cells. Cell Rep. 2019; 28.

  65. Leal F, Cohn MJ. Loss and re-emergence of legs in snakes by modular evolution of Sonic hedgehog and HOXD enhancers. Curr Biol. 2016;26:2966–73.

    Article  CAS  PubMed  Google Scholar 

  66. Barutcu AR, Lajoie BR, McCord RP, Tye CE, Hong D, Messier TL, Browne G, van Wijnen AJ, Lian JB, Stein JL, et al. Chromatin interaction analysis reveals changes in small chromosome and telomere clustering between epithelial and breast cancer cells. Genome Biol. 2015;16:214.

    Article  PubMed  PubMed Central  Google Scholar 

  67. Rowley MJ, Nichols MH, Lyu X, Ando-Kuri M, Rivera ISM, Hermetz K, Wang P, Ruan Y, Corces VG. Evolutionarily conserved principles predict 3D chromatin organization. Mol Cell. 2017;67:837.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Dong QL, Wang JB, Li XC, Gong L. Progresses in the plant 3D chromatin architecture. Yi Chuan Hereditas. 2020;42:73–86.

    PubMed  Google Scholar 

  69. Kloet SL, Karemaker ID, van Voorthuijsen L, Lindeboom RGH, Baltissen MP, Edupuganti RR, Poramba-Liyanage DW, Jansen PWTC, Vermeulen M. NuRD-interacting protein ZFP296 regulates genome-wide NuRD localization and differentiation of mouse embryonic stem cells. Nat Commun. 2018;9:4588.

    Article  PubMed  PubMed Central  Google Scholar 

  70. Lhoumaud P, Badri S, Rodriguez-Hernaez J, Sakellaropoulos T, Sethia G, Kloetgen A, Cornwell M, Bhattacharyya S, Ay F, Bonneau R, et al. NSD2 overexpression drives clustered chromatin and transcriptional changes in a subset of insulated domains. Nat Commun. 2019;10:4843.

    Article  PubMed  PubMed Central  Google Scholar 

  71. Dixon JR, Selvaraj S, Yue F, Kim A, Li Y, Shen Y, Hu M, Liu JS, Ren B. Topological domains in mammalian genomes identified by analysis of chromatin interactions. Nature. 2012;485:376–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Hou C, Li L, Qin ZS, Corces VG. Gene density, transcription, and insulators contribute to the partition of the Drosophila genome into physical domains. Mol Cell. 2012;48:471–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Hyle J, Zhang Y, Wright S, Xu B, Shao Y, Easton J, Tian L, Feng R, Xu P, Li C. Acute depletion of CTCF directly affects MYC regulation through loss of enhancer–promoter looping. Nucleic Acids Res. 2019;47:6699–713.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Van Bortle K, Nichols MH, Li L, Ong CT, Takenaka N, Qin ZS, Corces VG. Insulator function and topological domain border strength scale with architectural protein occupancy. Genome Biol. 2014;15:R82.

    Article  PubMed  PubMed Central  Google Scholar 

  75. Zhang H, Emerson DJ, Gilgenast TG, Titus KR, Lan Y, Huang P, Zhang D, Wang H, Keller CA, Giardine B, et al. Chromatin structure dynamics during the mitosis-to-G1 phase transition. Nature. 2019;576:158–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Cho SW, Xu J, Sun R, Mumbach MR, Carter AC, Chen YG, Yost KE, Kim J, He J, Nevins SA, et al. Promoter of lncRNA gene PVT1 is a tumor-suppressor DNA boundary element. Cell. 2018;173:1398.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Tameire F, Verginadis II, Leli NM, Polte C, Conn CS, Ojha R, Salas Salinas C, Chinga F, Monroy AM, Fu W, et al. ATF4 couples MYC-dependent translational activity to bioenergetic demands during tumour progression. Nat Cell Biol. 2019;21:889–99.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Gonen N, Futtner CR, Wood S, Garcia-Moreno SA, Salamone IM, Samson SC, Sekido R, Poulat F, Maatouk DM, Lovell-Badge R. Sex reversal following deletion of a single distal enhancer of Sox9. Science (New York, NY). 2018;360:1469–73.

    Article  CAS  Google Scholar 

  79. Tang Z, Luo OJ, Li X, Zheng M, Zhu JJ, Szalaj P, Trzaskoma P, Magalska A, Wlodarczyk J, Ruszczycki B, et al. CTCF-mediated human 3D genome architecture reveals chromatin topology for transcription. Cell. 2015;163:1611–27.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Wu H, Wu Y, Jiang Y, Zhou B, Zhou H, Chen Z, Xiong Y, Liu Q, Zhang H. scHiCStackL: a stacking ensemble learning-based method for single-cell Hi-C classification using cell embedding. Brief Bioinformatics. 2022; 23.

  81. Collombet S, Ranisavljevic N, Nagano T, Varnai C, Shisode T, Leung W, Piolot T, Galupa R, Borensztein M, Servant N, et al. Parental-to-embryo switch of chromosome organization in early embryogenesis. Nature. 2020;580:142–6.

    Article  CAS  PubMed  Google Scholar 

  82. Kottler MJ. From 48 to 46: cytological technique, preconception, and the counting of human chromosomes. Bull Hist Med. 1974;48:465–502.

    CAS  PubMed  Google Scholar 

  83. Young KH. Yeast two-hybrid: so many interactions, (in) so little time. Biol Reprod. 1998;58:302–11.

    Article  CAS  PubMed  Google Scholar 

  84. Chait BT. Mass spectrometry in the postgenomic era. Annu Rev Biochem. 2011;80:239–46.

    Article  CAS  PubMed  Google Scholar 

  85. Niu J, Zhang X, Li G, Yan P, Yan Q, Dai Q, Jin D, Shen X, Wang J, Zhang MQ, Gao J. A novel cytogenetic method to image chromatin interactions at subkilobase resolution: Tn5 transposase-based fluorescence in situ hybridization. J Genet Genomics. 2020;47:727–35.

    Article  PubMed  Google Scholar 

  86. Aljahani A, Hua P, Karpinska MA, Quililan K, Davies JOJ, Oudelaar AM. Analysis of sub-kilobase chromatin topology reveals nano-scale regulatory interactions with variable dependence on cohesin and CTCF. Nat Commun. 2022;13:2139.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Seidlitz T, Merker SR, Rothe A, Zakrzewski F, von Neubeck C, Grützmann K, Sommer U, Schweitzer C, Schölch S, Uhlemann H, et al. Human gastric cancer modelling using organoids. Gut. 2019;68:207–17.

    Article  CAS  PubMed  Google Scholar 

  88. Lott PC, Carvajal-Carmona LG. Resolving gastric cancer aetiology: an update in genetic predisposition. Lancet Gastroenterol Hepatol. 2018;3:874–83.

    Article  PubMed  PubMed Central  Google Scholar 

  89. Qiu Y, Huang S. CTCF-mediated genome organization and leukemogenesis. Leukemia. 2020;34:2295–304.

    Article  PubMed  PubMed Central  Google Scholar 

  90. Yan F, Powell DR, Curtis DJ, Wong NC. From reads to insight: a hitchhiker’s guide to ATAC-seq data analysis. Genome Biol. 2020;21:22.

    Article  PubMed  PubMed Central  Google Scholar 

  91. Cai Z, Cao C, Ji L, Ye R, Wang D, Xia C, Wang S, Du Z, Hu N, Yu X, et al. RIC-seq for global in situ profiling of RNA–RNA spatial interactions. Nature. 2020;582:432–7.

    Article  CAS  PubMed  Google Scholar 

  92. Johanson TM, Chan WF, Keenan CR, Allan RS. Genome organization in immune cells: unique challenges. Nat Rev Immunol. 2019;19:448–56.

    Article  CAS  PubMed  Google Scholar 

  93. Cao L, Liu S, Li Y, Yang G, Luo Y, Li S, Du H, Zhao Y, Wang D, Chen J, et al. The nuclear matrix protein SAFA surveils viral RNA and facilitates immunity by activating antiviral enhancers and super-enhancers. Cell Host Microbe. 2019;26:369.

    Article  CAS  PubMed  Google Scholar 

  94. Wang Y, Song F, Zhang B, Zhang L, Xu J, Kuang D, Li D, Choudhary MNK, Li Y, Hu M, et al. The 3D Genome Browser: a web-based browser for visualizing 3D genome organization and long-range chromatin interactions. Genome Biol. 2018;19:151.

    Article  PubMed  PubMed Central  Google Scholar 

  95. Haigh JL, Adhikari A, Copping NA, Stradleigh T, Wade AA, Catta-Preta R, Su-Feher L, Zdilar I, Morse S, Fenton TA, et al. Deletion of a non-canonical regulatory sequence causes loss of Scn1a expression and epileptic phenotypes in mice. Genome Med. 2021;13:69.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Cao X, Xing L, He H, Zhang X. Views on GWAS statistical analysis. Bioinformation. 2020;16:393–7.

    PubMed  PubMed Central  Google Scholar 

  97. Hawe JS, Wilson R, Schmid KT, Zhou L, Lakshmanan LN, Lehne BC, Kühnel B, Scott WR, Wielscher M, Yew YW, et al. Genetic variation influencing DNA methylation provides insights into molecular mechanisms regulating genomic function. Nat Genet. 2022;54:18–29.

    Article  CAS  PubMed  Google Scholar 

  98. Huang K-L, Marcora E, Pimenova AA, Di Narzo AF, Kapoor M, Jin SC, Harari O, Bertelsen S, Fairfax BP, Czajkowski J, et al. A common haplotype lowers PU.1 expression in myeloid cells and delays onset of Alzheimer’s disease. Nat Neurosci. 2017;20:1052–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Peterson RE, Kuchenbaecker K, Walters RK, Chen CY, Popejoy AB, Periyasamy S, Lam M, Iyegbe C, Strawbridge RJ, Brick L, et al. Genome-wide association studies in ancestrally diverse populations: opportunities, methods, pitfalls, and recommendations. Cell. 2019;179:589–603.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Dong X, Liu C, Dozmorov M. Review of multi-omics data resources and integrative analysis for human brain disorders. Brief Funct Genomics. 2021;20:223–34.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Chen X, Xu H, Yuan P, Fang F, Huss M, Vega VB, Wong E, Orlov YL, Zhang W, Jiang J, et al. Integration of external signaling pathways with the core transcriptional network in embryonic stem cells. Cell. 2008;133:1106–17.

    Article  CAS  PubMed  Google Scholar 

  102. Deng W, Lee J, Wang H, Miller J, Reik A, Gregory PD, Dean A, Blobel GA. Controlling long-range genomic interactions at a native locus by targeted tethering of a looping factor. Cell. 2012;149:1233–44.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Krivega I, Dean A. LDB1-mediated enhancer looping can be established independent of mediator and cohesin. Nucleic Acids Res. 2017;45:8255–68.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Zheng H, Xie W. The role of 3D genome organization in development and cell differentiation. Nat Rev Mol Cell Biol. 2019;20:535–50.

    Article  CAS  PubMed  Google Scholar 

  105. Shahini A, Choudhury D, Asmani M, Zhao R, Lei P, Andreadis ST. NANOG restores the impaired myogenic differentiation potential of skeletal myoblasts after multiple population doublings. Stem Cell Res. 2018;26:55–66.

    Article  CAS  PubMed  Google Scholar 

  106. Stevanovic M, Drakulic D, Lazic A, Ninkovic DS, Schwirtlich M, Mojsin M. SOX transcription factors as important regulators of neuronal and glial differentiation during nervous system development and adult neurogenesis. Front Mol Neurosci. 2021;14: 654031.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Liu X, Peng L, Li D, He C, Xing S, Wang Y, He Y. The impacts of and genetic variants on rheumatoid arthritis risk in the Chinese Han population: a case-control study. Int J General Med. 2021;14:2147–59.

    Article  Google Scholar 

  108. Shi Y, Norberg E, Vakifahmetoglu-Norberg H. Mutant p53 as a regulator and target of autophagy. Front Oncol. 2020;10: 607149.

    Article  PubMed  Google Scholar 

  109. Mallo M. Reassessing the role of hox genes during vertebrate development and evolution. Trends Genet TIG. 2018;34:209–17.

    Article  CAS  PubMed  Google Scholar 

  110. Blow MJ, McCulley DJ, Li Z, Zhang T, Akiyama JA, Holt A, Plajzer-Frick I, Shoukry M, Wright C, Chen F, et al. ChIP-Seq identification of weakly conserved heart enhancers. Nat Genet. 2010;42:806–10.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Alam H, Tang M, Maitituoheti M, Dhar SS, Kumar M, Han CY, Ambati CR, Amin SB, Gu B, Chen T-Y, et al. KMT2D deficiency impairs super-enhancers to confer a glycolytic vulnerability in lung cancer. Cancer Cell. 2020;37:599.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Whyte WA, Orlando DA, Hnisz D, Abraham BJ, Lin CY, Kagey MH, Rahl PB, Lee TI, Young RA. Master transcription factors and mediator establish super-enhancers at key cell identity genes. Cell. 2013;153:307–19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Hnisz D, Abraham BJ, Lee TI, Lau A, Saint-André V, Sigova AA, Hoke HA, Young RA. Super-enhancers in the control of cell identity and disease. Cell. 2013;155:934–47.

    Article  CAS  PubMed  Google Scholar 

  114. Hnisz D, Schuijers J, Lin CY, Weintraub AS, Abraham BJ, Lee TI, Bradner JE, Young RA. Convergence of developmental and oncogenic signaling pathways at transcriptional super-enhancers. Mol Cell. 2015;58:362–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Dowen JM, Fan ZP, Hnisz D, Ren G, Abraham BJ, Zhang LN, Weintraub AS, Schujiers J, Lee TI, Zhao K, Young RA. Control of cell identity genes occurs in insulated neighborhoods in mammalian chromosomes. Cell. 2014;159:374–87.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Chen H, Liang H. A high-resolution map of human enhancer RNA loci characterizes super-enhancer activities in cancer. Cancer Cell. 2020;38:701.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Pott S, Lieb JD. What are super-enhancers? Nat Genet. 2015;47:8.

    Article  CAS  PubMed  Google Scholar 

  118. Zhang J, Yue W, Zhou Y, Liao M, Chen X, Hua J. Super enhancers—functional cores under the 3D genome. Cell Prolif. 2021;54: e12970.

    Article  CAS  PubMed  Google Scholar 

  119. Lee HK, Willi M, Kuhns T, Liu C, Hennighausen L. Redundant and non-redundant cytokine-activated enhancers control Csn1s2b expression in the lactating mouse mammary gland. Nat Commun. 2021;12:2239.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Xu L, Chen Y, Huang Y, Sandanaraj E, Yu JS, Lin RY-T, Dakle P, Ke X-Y, Chong YK, Koh L, et al. Topography of transcriptionally active chromatin in glioblastoma. Sci Adv. 2021; 7.

  121. Kokkinos J, Jensen A, Sharbeen G, McCarroll JA, Goldstein D, Haghighi KS, Phillips PA. Does the microenvironment hold the hidden key for functional precision medicine in pancreatic cancer? Cancers. 2021;13:2427.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Hagihara Y, Yoshimatsu Y, Mikami Y, Takada Y, Mizuno S, Kanai T. Epigenetic regulation of T helper cells and intestinal pathogenicity. Semin Immunopathol. 2019;41:379–99.

    Article  PubMed  Google Scholar 

  123. Liu S, Zhao K. The toolbox for untangling chromosome architecture in immune cells. Front Immunol. 2021;12: 670884.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Tannock IF, Hickman JA. Limits to precision cancer medicine. N Engl J Med. 2017;376:96–7.

    PubMed  Google Scholar 

  125. Abrahams E, Eck SL. Molecular medicine: precision oncology is not an illusion. Nature. 2016;539:357.

    Article  CAS  PubMed  Google Scholar 

  126. Letai A. Functional precision cancer medicine-moving beyond pure genomics. Nat Med. 2017;23:1028–35.

    Article  CAS  PubMed  Google Scholar 

  127. Lynn RC, Weber EW, Sotillo E, Gennert D, Xu P, Good Z, Anbunathan H, Lattin J, Jones R, Tieu V, et al. c-Jun overexpression in CAR T cells induces exhaustion resistance. Nature. 2019;576:293–300.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Ou HD, Phan S, Deerinck TJ, Thor A, Ellisman MH, O'Shea CC. ChromEMT: Visualizing 3D chromatin structure and compaction in interphase and mitotic cells. Science (New York, NY). 2017; 357.

  129. Liu X, Zhang Y, Chen Y, Li M, Zhou F, Li K, Cao H, Ni M, Liu Y, Gu Z, et al. In situ capture of chromatin interactions by biotinylated dCas9. Cell. 2017;170:1028.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Navarro Gonzalez J, Zweig AS, Speir ML, Schmelter D, Rosenbloom KR, Raney BJ, Powell CC, Nassar LR, Maulding ND, Lee CM, et al. The UCSC genome browser database: 2021 update. Nucleic Acids Res. 2021;49:D1046–57.

    Article  PubMed  Google Scholar 

  131. Robinson JT, Turner D, Durand NC, Thorvaldsdóttir H, Mesirov JP, Aiden EL. Juicebox.js provides a cloud-based visualization system for Hi-C data. Cell Syst. 2018; 6.

  132. Yang D, Jang I, Choi J, Kim MS, Lee AJ, Kim H, Eom J, Kim D, Jung I, Lee B. 3DIV: a 3D-genome interaction viewer and database. Nucleic Acids Res. 2018;46:D52-d57.

    Article  CAS  PubMed  Google Scholar 

  133. Li D, Hsu S, Purushotham D, Sears RL, Wang T. WashU epigenome browser update 2019. Nucleic Acids Res. 2019;47:W158–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Dekker J, Rippe K, Dekker M, Kleckner N. Capturing chromosome conformation. Science (New York, NY). 2002;295:1306–11.

    Article  CAS  Google Scholar 

  135. Simonis M, Klous P, Splinter E, Moshkin Y, Willemsen R, de Wit E, van Steensel B, de Laat W. Nuclear organization of active and inactive chromatin domains uncovered by chromosome conformation capture-on-chip (4C). Nat Genet. 2006;38:1348–54.

    Article  CAS  PubMed  Google Scholar 

  136. Ruan Y. Genome-sequencing anniversary. Presenting the human genome: now in 3D! Science (New York, NY). 2011;331:1025–6.

    Article  Google Scholar 

  137. Fang R, Yu M, Li G, Chee S, Liu T, Schmitt AD, Ren B. Mapping of long-range chromatin interactions by proximity ligation-assisted ChIP-seq. Cell Res. 2016;26:1345–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. Belaghzal H, Dekker J, Gibcus JH. Hi-C 2.0: an optimized Hi-C procedure for high-resolution genome-wide mapping of chromosome conformation. Methods (San Diego, Calif). 2017;123:56–65.

    Article  CAS  PubMed  Google Scholar 

  139. Zhang H, Qin Z, Yue X, Liu Y, Sun X, Feng J, Xu Z, Zhao J, Li K, Qiu J, et al. Proteome-wide profiling of transcriptional machinery on accessible chromatin with biotinylated transposons. Sci Adv. 2021;7:eabh1022.

    Article  CAS  PubMed  Google Scholar 

  140. Chen X, Shen Y, Draper W, Buenrostro JD, Litzenburger U, Cho SW, Satpathy AT, Carter AC, Ghosh RP, East-Seletsky A, et al. ATAC-see reveals the accessible genome by transposase-mediated imaging and sequencing. Nat Methods. 2016;13:1013–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Nagano T, Lubling Y, Yaffe E, Wingett SW, Dean W, Tanay A, Fraser P. Single-cell Hi-C for genome-wide detection of chromatin interactions that occur simultaneously in a single cell. Nat Protoc. 2015;10:1986–2003.

    Article  CAS  PubMed  Google Scholar 

  142. Ramani V, Deng X, Qiu R, Gunderson KL, Steemers FJ, Disteche CM, Noble WS, Duan Z, Shendure J. Massively multiplex single-cell Hi-C. Nat Methods. 2017;14:263–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank Dr. Dev Sooranna for his excellent language editing work for this manuscript.

Funding

This work was funded by grants from the National Natural Science Foundation of China (32102514, U20A2051, and31972996), the Guangxi Natural Science Foundation (2020GXNSFBA297148 and AD20159062), the State Key Laboratory for Conservation and Utilization of Subtropical Agro-Bioresources (SKLCUSA-a202203), the Guangxi Distinguished Scholars Program (201835), and the Bama County Program for Talents in Science and Technology, Guangxi, China (202101265 and 202100174).

Author information

Authors and Affiliations

Authors

Contributions

M.C. and X.L.: conceptualization, methodology, visualization, writing—original draft. Q.L. and H.L: writing—review and editing. H.L. and D.S.: conceptualization, methodology, supervision, writing—review and editing. All the authors read and approved the final manuscript.

Corresponding authors

Correspondence to Deshun Shi or Hui Li.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chen, M., Liu, X., Liu, Q. et al. 3D genomics and its applications in precision medicine. Cell Mol Biol Lett 28, 19 (2023). https://doi.org/10.1186/s11658-023-00428-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s11658-023-00428-x

Keywords