Next Article in Journal
Genomic Landscape of Myelodysplastic/Myeloproliferative Neoplasms: A Multi-Central Study
Previous Article in Journal
Needle and Branch Trait Variation Analysis and Associated SNP Loci Mining in Larix olgensis
Previous Article in Special Issue
Mathematical Modeling and Inference of Epidermal Growth Factor-Induced Mitogen-Activated Protein Kinase Cell Signaling Pathways
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Polymer Physics Models Reveal Structural Folding Features of Single-Molecule Gene Chromatin Conformations

1
Dipartimento di Fisica, Università di Napoli Federico II, and INFN Napoli, Complesso Universitario di Monte Sant’Angelo, 80126 Naples, Italy
2
Department of Systems Biology, University of Massachusetts Chan Medical School, Worcester, MA 01605, USA
3
Diabetes Center of Excellence and Program in Molecular Medicine, University of Massachusetts Chan Medical School, Worcester, MA 01655, USA
4
DIETI, Università di Napoli Federico II, Via Claudio 21, 80125 Naples, Italy
5
INFN Napoli, Complesso Universitario di Monte Sant’Angelo, 80126 Naples, Italy
6
Howard Hughes Medical Institute, Chevy Chase, MD 20815, USA
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2024, 25(18), 10215; https://doi.org/10.3390/ijms251810215
Submission received: 14 July 2024 / Revised: 17 September 2024 / Accepted: 22 September 2024 / Published: 23 September 2024

Abstract

:
Here, we employ polymer physics models of chromatin to investigate the 3D folding of a 2 Mb wide genomic region encompassing the human LTN1 gene, a crucial DNA locus involved in key cellular functions. Through extensive Molecular Dynamics simulations, we reconstruct in silico the ensemble of single-molecule LTN1 3D structures, which we benchmark against recent in situ Hi-C 2.0 data. The model-derived single molecules are then used to predict structural folding features at the single-cell level, providing testable predictions for super-resolution microscopy experiments.

1. Introduction

Recent DNA technologies, such as Hi-C [1,2], GAM [3,4], and SPRITE [5,6], have shown that mammalian chromosomes have complex, non-random 3D architectures within the cell nucleus, encompassing multiple folding structures across genomic scales [7,8,9,10,11,12,13]. Such an organization includes, for example, DNA loops [14], Topologically Associating Domains (TADs) [15,16], metaTADs [17], and larger architectural features, such as A/B compartments [1], lamina-associated interactions [18], and nuclear chromosome territories [19]. These chromatin structures play critical roles in gene regulation, as distal DNA regulatory elements (e.g., enhancers) have been reported to establish specific physical contacts with their target genes, particularly within TADs, to orchestrate cell transcriptional programs [20,21,22,23,24,25,26,27,28,29,30,31]. Misfolding of chromatin 3D structure, which can lead to abnormal gene–enhancer interactions, has been indeed implicated in genetic disease [32,33,34,35,36,37,38,39,40,41,42,43].
Moreover, advancements in super-resolution microscopy techniques have enabled the direct visualization of chromatin conformations at nanometer scales within individual cell nuclei, providing quantitative insights into single-cell genome structures [44]. These methods have revealed, for instance, that TAD-like domains exhibit spatially segregated globular conformations in single cells, albeit with significant cell-to-cell variability, highlighting crucial constraints on chromatin folding at the level of single DNA molecules [45,46,47,48,49,50,51,52,53,54,55,56,57].
On the other hand, theoretical models of chromatin have been instrumental in bolstering experimental technologies to understand the 3D organization of the genome [58]. These models, including polymer physics-based and computational approaches [59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112], have been used, e.g., to dissect the fundamental molecular mechanisms shaping DNA contact formation and to derive quantitatively testable predictions, particularly in disease contexts [113,114,115,116,117,118,119,120,121,122,123,124].
In this work, we focus on a reference polymer physics approach, the Strings and Binders (SBS) model of chromatin [86,87], which has been extensively validated against independent experimental datasets, such as bulk Hi-C, GAM, and single-cell microscopy data [125,126,127,128]. In the SBS picture, physical contacts between distal DNA sites (such as genes and enhancers) are established by diffusing cognate binding factors, which can bridge DNA binding sites through thermodynamic mechanisms of phase transitions [61,127]. Here, we apply the SBS model to investigate the single-molecule folding of a 2 Mb wide genomic locus (Chr21:28–30 Mb, hg38) encompassing the LTN1 gene in human WTC-11 cells (a human-induced pluripotent stem cell line), a crucial locus involved in diverse cellular functions ranging, e.g., from embryonic development and targeting of misfolded proteins to the onset of neurodegenerative diseases [129,130,131,132,133,134,135,136].
By performing massive Molecular Dynamics (MD) simulations of the model, we derive an ensemble of in silico 3D conformation of the gene locus that we validate against recent in situ Hi-C 2.0 data generated via a highly optimized Hi-C protocol [137] available from the 4DN Data Portal (dataset reference number 4DNESJ7S5NDJ) [138,139]. Next, by leveraging the model single-polymer conformations, we conduct structural analyses at the single-molecule level, including spatial distance matrices, assessment of cell-to-cell folding variability, and 3D shape descriptors, producing non-trivial predictions that can be tested by real single-cell microscopy (e.g., super-resolution multiplexed FISH) experiments.
Overall, the model provides a validated, quantitative blueprint for assessing the spatial organization of key human genomic loci at the single-molecule level, complementing empirical investigations in the understanding of chromatin structures in single cells.

2. Methods

2.1. The Strings and Binders (SBS) Polymer Model

The Strings and Binders (SBS) polymer model envisages a theoretical framework for understanding the 3D organization of chromatin where molecular interactions between distant genomic regions are driven by diffusing agents, such as Transcription Factors (TFs) or coactivators, that diffuse in the nuclear environment [86,87]. In the model, a chromosomal segment is represented as a coarse-grained, self-avoiding walk (SAW) polymer chain with specific binding sites for molecular bridging binders (a schematic model cartoon is shown in Figure 1a).
In the simplest case of an SBS homopolymer chain, where the binding sites are all equal, the model has a phase transition from an extended coil (randomly folded) to a compact globular state [140] as soon as the binder molar concentration or its attractive interaction strength exceeds critical thresholds (see also below). For weak biochemical affinities (i.e., units of KBT) and for genomic resolutions approaching the sub-megabase scale, these thresholds are in the range of a few micromole/l [61,127], consistent with typical TF concentrations observed in vitro [141].
To dissect the folding dynamics of real chromatin regions characterized by complex genomic contacts, the SBS model can be extended to include multiple types of binding sites, each associated with a specific type of cognate molecular binder [61,142]. Such heteropolymer configurations induce micro-phase separations of the polymer chain, leading to the formation of distinct globular domains enriched with specific binding motifs [127]. The features of these binding sites (i.e., their genomic location and relative abundance) are determined through a previously published machine learning approach, named PRISMR [117], which employs standard Monte Carlo-based optimization procedures to infer the optimal (i.e., minimal) SBS model of specific chromatin regions based solely on input bulk (e.g., Hi-C) contact data, with no additional fitting parameters. In our studied LTN1 locus in human WTC-11 cells (Chr21: 28–30 Mb), PRISMR returns a polymer of 800 beads with ten distinct types of binding sites, visually represented by different colors (Figure 1a). Interestingly, these domains have an overlapping, combinatorial organization along the chain, which has been shown to be crucial in explaining chromatin contacts with high genome-wide accuracy [143].

2.2. Molecular Dynamics (MD) Simulations

To derive a statistical ensemble of single-molecule model 3D structures of the LTN1 locus, we performed extensive Molecular Dynamics (MD) simulations within the freely available LAMMPS software (version 30 July 2016) [144], optimized for parallel computation [145,146]. In the model MD implementation, the polymer is a standard coarse-grained bead-spring chain [147], and the binders are simple spherical particles. The motion of polymer beads and binders is determined by a Langevin equation with standard parameters (friction coefficient ζ = 0.5 and temperature T = 1, dimensionless units), numerically integrated via the Velocity–Verlet algorithm [148]. For simplicity, in our simulations, we set the diameter of both polymer beads and binders to σ = 1, using it as the unit length of the model; similarly, we set the mass of beads and binders to be equal, taking it as the reference mass unit m = 1 [147].
The physical interaction potentials are set as in classical polymer simulation studies [147]. Excluded volume interactions between consecutive beads are modeled using a truncated, purely repulsive Lennard–Jones (LJ) potential, adjacent polymer beads are connected by FENE bonds with standard parameters (maximum length 1.6σ and spring constant 30KBT/σ2), and attractive interactions between beads and cognate binders are described by a short-range, truncated LJ (cut-off distance 1.5σ).
The initial states of the MD simulations are open SAW configurations, located in a cubic simulation box of size 50σ with periodic boundary conditions to control finite size effects [149]. Binders are randomly injected into the simulation box, and then the system of beads and binders is equilibrated for 108 MD time iteration steps. To monitor the system folding dynamics, we recorded the time track of the polymer gyration radius, Rg. This function has a sharp drop at a characteristic time scale and then plateaus, marking the phase transition of the polymer from an initial free SAW chain to an equilibrium phase-separated globular conformation (Figure 1b). To give a sense of the chain compaction, the linear size of the polymer decreases by around 70% in the transition [140], resulting in an average Rg in the phase-separated state equal to 6.4σ.
To test model robustness, we sampled a broad range of bead-binder affinity values in the weak biochemical energy scale, from 0 to 8 KBT, and explored up to three orders of magnitude in binder concentration as detailed in previous studies [61,127,150]. As explained above, the number of binders (or their energy affinities) serves as a key control parameter in the model folding, sharply driving its switch-like conformational change from the initial coil to the equilibrium phase-separated globular state, consistent with classic results of polymer phase transitions (Figure 2) [140]. For each set of parameters, we performed up to 3 × 102 independent runs to ensure statistical strength.
The MD length scale (i.e., the bead diameter σ) can be mapped into physical units using the relation [87] σ = (s/G)D1/3, where s is the genomic content per bead (s = L/N, where L = 2 Mb is the genomic length of the locus and N = 800 the number of beads), G is the genome length (6 Gb), and D is the cell nucleus diameter (taken to be 10 μm as an order of magnitude). These approximations yield σ = 74.7 nm, which is in line with previous polymer physics studies [77,83]. As an additional check, single-cell microscopy data of the same LTN1 genomic region in human IMR90 cells reported an average Rg of the locus equal to 464 nm [45]. This value, matched with that found in our simulations (6.4σ), returns σ = 72.5 nm, which is remarkably close to our previous calculation. In the following, to set a reference value for σ, we take the average of those two independent estimates, i.e., σ = 73.6 nm.
Finally, the simulation time, τ, can be converted into physical time via the formula τ =ησ3/(KBT), where η is the nucleoplasm viscosity. Given typical viscosity values in the range of a few fractions of poise [63,151], the model time scale is on the order of milliseconds, which is consistent with classic chromatin simulations [63].

3. Results

3.1. Folding of the SBS Model of the Human LTN1 Gene Locus

To comprehensively investigate the folding properties of our model at the ensemble population level, we measured the average pairwise contact matrix in its phase-separated state. This was achieved by computing the mean of the model single-molecule contact maps, i.e., symmetric square matrices where each entry, Aij, is either 1 or 0, depending on whether the polymer sites i and j are in contact. A contact event is considered to occur if the spatial distance between the sites is below a typical distance threshold [63,117]. For our study, we explored thresholds within the range from 2 up to 5σ, corresponding to a spatial distance range of about 150–350 nm (see above), and they all provided analogous results. This range is consistent with established reference contact threshold values from microscopy studies [51].
To benchmark and validate the model output, we used publicly available in situ Hi-C 2.0 data [137] of the LTN1 locus in human WTC-11 cells from the 4DN Data Portal (dataset reference number 4DNESJ7S5NDJ) [138]. The experimental Hi-C contact map, which is binned at a 25 kb resolution to match the model coarse-graining level, exhibits specific and non-random patterns of contacts, including the presence of TADs and sub-TAD domains, inter-TAD interactions, and long-range (>500 kb) looping contacts, particularly around the LTN1 gene (Figure 3a, left panel).
Remarkably, the contact map generated by the model captures these observed experimental features, successfully recapitulating the overall structure of pairwise interactions (Figure 3a, right). This agreement is quantitatively supported by the high Pearson correlation coefficient between the model and Hi-C contact matrices, r = 0.90. As an additional measure of similarity, we computed the genomic distance corrected Pearson correlation coefficient, r’, which averages out trivial genomic proximity effects [117]. Despite the minimal ingredients of the model, we find r′ = 0.59, which is a comparatively high correlation value considering that a randomly folded control chain would produce an r’ close to zero [152].
To further validate model robustness, we also calculated the Spearman correlation coefficient, rs, between the model and experimental contact maps. Our analysis yields a high rs value of 0.76, indicating a strong statistical similarity [97,153].
Finally, to check the performance of the model at different genomic length scales, we performed a more local analysis by computing the contact probability, Pc(s), of the model and Hi-C 2.0 data. Our analysis confirms that the model closely aligns with experimental data across three orders of magnitude of genomic separations (Figure 3b; Pearson corr. between the curves r = 0.98).
To summarize, the ensemble of model conformations of the LTN1 locus aligns well with Hi-C experimental data, demonstrating that our model quantitatively captures in silico the structural features of the locus folding at the population-average level.

3.2. Structural Heterogeneity of LTN1 Single-Molecule Conformations in the Model

Next, we aimed to study the folding dynamics of the locus as predicted by our simulations at the single-molecule level. By leveraging the 3D coordinates generated through the SBS model, we computed spatial distance maps for each phase-separated polymer conformation. These maps are symmetric square matrices of pairwise Euclidean distances between polymer sites across the locus. Our analysis reveals distinct single-molecule distance patterns, featuring TAD-like domains occurring at different genomic positions and long-range loop contacts spanning across TAD boundaries (Figure 4a). Consistent with previous studies [45,53,154,155], this variability underscores the highly dynamic nature of the 3D structures of single chromatin conformations. In the SBS model, such a structural heterogeneity arises, beyond stochastic thermal fluctuations, from the inherent folding degeneracy due to the specific, overlapping distribution of its binding sites (see, e.g., Figure 1a) [127].
To quantify the extent of structural variability among model single molecules, we measured the r’ pairwise correlations between their distance matrices. In the absence of a correlation between matrices, we would expect an r’ distribution centered around zero. Conversely, perfectly correlated matrices would yield an r’ distribution peaked at 1. Interestingly, our calculations return an unimodal distribution with a mean value of 0.13 and a variance of 0.15 (Figure 4b), which suggests a substantial heterogeneity of polymer 3D structures that, though, retain a residual degree of structural correlation (as indicated by the non-zero r’ average value). These results are in line with recent microscopy studies that reported a strong cell-to-cell variability of individual chromatin conformations at the sub-megabase scale [45,46,52], revealing, nevertheless, sub-clusters of structures with correlated behaviors [51].
To summarize, the model highlights a broad distribution of single-molecule 3D structures at the human LTN1 locus, providing quantitative predictions (e.g., single-molecule spatial distances and all-against-all pair correlations), which can be directly tested in real biological contexts by independent single-cell microscopy approaches, including, for instance, super-resolution multiplexed FISH or 3D-SIM-based techniques [156,157,158]. Such experimental validation would not only assess the predictive power of the model but also deepen our understanding of the intrinsic structural diversity and dynamic behavior of genomic loci at the single-molecule level.

3.3. Shape and Size of 3D Model Single Molecules

To further characterize the LTN1 single-molecule structures predicted by our chromatin model, we calculated their shape and volume. Specifically, for each model conformation in the phase-separated state, we computed the inertia tensor, defined as I β γ = α = 1 N m α r α 2 δ β γ r α β r α γ , where N is the number of monomers along the polymer chain, mα is the mass of the α-th monomer, rαβ is the β-th spatial coordinate, and β,γ are the spatial component indexes equal to {0,1,2}. The eigenvalues of this tensor, corresponding to the system principal moments of inertias Ia, Ib, and Ic, can be related through standard textbook formulas to the semi-axes (a, b, c) of a triaxial ellipsoid enclosing the volume contour of a given conformation as follows: a 2 = 5 2 N I b I a + I c , b 2 = 5 2 N I a I b + I c , and c 2 = 5 2 N ( I a I c + I b ) .
In the case of a perfectly spherical conformation, the three eigenvalues are equal, and that implies a = b = c. However, in the case of the LTN1 locus, our model predicts substantial deviations of individual chromatin configurations from a spherical topology, as they are found to have a prolate shape with a > bc (Figure 5a). Indeed, while the distribution of the ratios b/c has an average value of 1.0 (orange histogram in the figure), the ratios a/c are centered around 2.0 (blue histogram); moreover, statistical analysis indicates that these two distributions are significantly distinguishable from each other (two-sided Mann–Whitney p-value < 0.001).
Using the inertia tensor, we also computed the volume V of single-molecule structures via the formula V = 4/3πabc, which returns an average volume estimation of the entire locus equal to 1500σ3. To give a sense of the physical length scales, by taking σ = 73.6 nm (see above), the model predicts an average locus volume of approximately 0.59 μm³. We also considered an alternative volume calculation using the formula V = 4/3πRg3, which returned a mean LTN1 volume equal to 1300σ3 (i.e., 0.52 μm³), comparable to our previous estimate.
Overall, these findings are in line with analogous volume measurements of Mb-wide chromatin loci, which typically fall within the range of fractions of a few μm3, as measured, e.g., by super-resolution microscopy and DNAseqFISH+ experiments across different species and cell types [45,51,54].
As an additional test on the shape of model-predicted conformations, we computed the polymer gyration tensor, defined as [159] T β γ = 1 N α = 1 N m α r α β r ¯ β r α γ r ¯ γ , where rαβ (rαγ) is the β-th (γ-th) component of the position vector of the α-th monomer and r ¯ β ( r ¯ γ ) is the component of the mass center of the polymer chain along the β-th (γ-th) direction. By diagonalizing the tensor, we derived its three eigenvalues, λ1 ≥ λ2 ≥ λ3, from which we calculated the ellipticity, ε, of the model single molecules, defined as ε = 2λ3/(λ1 + λ2). For spherical configurations, ε = 1, as the three eigenvalues are expected to be equal [159]. Consistent with the observed high structural variability of the LTN1 polymer conformations, we find that the distribution of single-molecule ellipticity values is broad (variance = 0.16), yet it has an average of 0.51, well below that of a spherical control (Figure 5b), highlighting the prolate nature of the single-molecule structures predicted by the model. Targeted microscopy experiments, specifically designed at the LTN1 region in WTC-11 cells, would provide a direct validation of those structural predictions.
Taken together, these results indicate that the LTN1 locus in human WTC-11 cells significantly deviates from a simple spherical geometry, predominantly displaying a prolate shape with substantial single-molecule structural variability, as predicted by our chromatin model and tested through multiple, independent shape and volume metrics.

4. Discussion

In this work, we combined polymer physics models of chromatin and computer simulations to comprehensively analyze the folding properties of the LTN1 gene locus in a human-induced pluripotent stem cell line. By computing the average pairwise contact matrix of the model, we demonstrated that its predicted patterns are consistent with available in situ Hi-C 2.0 experiments [137,138], capturing with high accuracy the complex architecture of TADs, sub-TADs, and long-range looping interactions observed at the gene locus.
Beyond population averages, our model reveals significant structural heterogeneity among individual chromatin 3D structures. This variability is evidenced, for instance, by distinct folding patterns in single-molecule spatial distance matrices and by the broad distribution of their pairwise correlation values. Our results suggest that while the LTN1 locus exhibits a substantial range of conformations, it also maintains a degree of structural correlation across the ensemble of single molecules, which is consistent with observations from recent microscopy studies [45,51].
Furthermore, our calculations of the inertia and gyration tensors provide quantitative insights into the shape and volume of the locus conformations. The model predicts a prolate shape of single molecules, with an average volume estimate in line with super-resolution imaging data [45,54]. It also provides additional testable predictions, such as the distribution of single-molecule ellipticity values, which can be directly validated by advanced microscopy techniques.
Albeit simplified, the model is robust as it is dictated by thermodynamics, and phase transitions and complex emergent behaviors are common to biological and soft-matter physics systems [160,161,162,163,164,165,166,167,168,169,170,171,172,173,174,175,176,177,178].
These findings have significant implications beyond structural characterization. Understanding the variability and dynamics of genomic loci at the single-molecule level is crucial for elucidating the principles of gene regulation, chromatin organization, and cellular function. The observed heterogeneity in 3D structures, indeed, could influence gene expression patterns, regulatory element interactions, and overall genomic stability.
The ability to predict and verify the structural configurations of genomic loci through principled computational models enhances our capacity to interpret how physical changes in individual chromatin structures can lead to functional outcomes, such as differential gene expression, cellular differentiation, and genetic-disease-associated phenotypes. For instance, the model can be used to detect ectopic gene–enhancer interactions, resulting, e.g., from genomic structural variants (SVs) at transcribing loci, which can contribute to developmental disease phenotypes [117,179,180]. These structural predictions can be experimentally validated through independent Hi-C and single-cell microscopy studies using, for example, patient-derived samples, thus offering molecular insights into disease mechanisms that could guide the development of targeted therapies.
This provides a robust basis for experimental validation at the single-cell level and a deeper understanding of the dynamic and heterogeneous nature of chromatin 3D architectures.

Author Contributions

Conceptualization, M.C., A.E. and M.N.; writing—original draft preparation, M.C. and A.A.; writing—review and editing, M.C., A.A. and A.E.; formal analysis, M.C., A.A. and A.E.; data curation and production, L.Y., J.H.G., K.M.P. and J.D.; visualization, M.C., A.A., A.E., F.V., A.F. and F.D.P.; supervision, M.C. and M.N. All authors have read and agreed to the published version of the manuscript.

Funding

MN acknowledges support from the National Institutes of Health Common Fund 4D Nucleome Program grant 5 1UM1HG011585-03, NextGeneration EU PNRR MUR M4C2 CN00000041 “National Center for Gene Therapy and Drugs based on RNA Technology” CUP E63C22000940007, MUR PRIN 2022 2022R8YXMR CUP E53D2300181 0006, MUR PRIN PNRR 2022 CUP E53D23018360001. J.D. acknowledges support by a grant from the National Institutes of Health Common Fund 4D Nucleome Program (U54-DK107980, UM1-HG011536).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request. Hi-C 2.0 data used in this study are available at the 4DN Data Portal [138] (dataset reference number 4DNESJ7S5NDJ).

Acknowledgments

We acknowledge computer resources from INFN, CINECA, ENEA CRESCO/ENEAGRID [181], and Scope/ReCAS/Ibisco at the University of Naples. J.D. is an investigator at the Howard Hughes Medical Institute.

Conflicts of Interest

J.D. is a member of the scientific advisory board of Arima Genomics, San Diego, CA, USA, and Omega Therapeutic, Cambridge, MA, USA.

References

  1. Lieberman-Aiden, E.; van Berkum, N.L.; Williams, L.; Imakaev, M.; Ragoczy, T.; Telling, A.; Amit, I.; Lajoie, B.R.; Sabo, P.J.; Dorschner, M.O.; et al. Comprehensive Mapping of Long-Range Interactions Reveals Folding Principles of the Human Genome. Science 2009, 326, 289–293. [Google Scholar] [CrossRef] [PubMed]
  2. Belton, J.M.; McCord, R.P.; Gibcus, J.H.; Naumova, N.; Zhan, Y.; Dekker, J. Hi-C: A Comprehensive Technique to Capture the Conformation of Genomes. Methods 2012, 58, 268–276. [Google Scholar] [CrossRef] [PubMed]
  3. Beagrie, R.A.; Scialdone, A.; Schueler, M.; Kraemer, D.C.A.; Chotalia, M.; Xie, S.Q.; Barbieri, M.; De Santiago, I.; Lavitas, L.M.; Branco, M.R.; et al. Complex Multi-Enhancer Contacts Captured by Genome Architecture Mapping. Nature 2017, 543, 519–524. [Google Scholar] [CrossRef]
  4. Winick-Ng, W.; Kukalev, A.; Harabula, I.; Zea-Redondo, L.; Szabó, D.; Meijer, M.; Serebreni, L.; Zhang, Y.; Bianco, S.; Chiariello, A.M.; et al. Cell-Type Specialization Is Encoded by Specific Chromatin Topologies. Nature 2021, 599, 684–691. [Google Scholar] [CrossRef] [PubMed]
  5. Quinodoz, S.A.; Ollikainen, N.; Tabak, B.; Palla, A.; Schmidt, J.M.; Detmar, E.; Lai, M.M.; Shishkin, A.A.; Bhat, P.; Takei, Y.; et al. Higher-Order Inter-Chromosomal Hubs Shape 3D Genome Organization in the Nucleus. Cell 2018, 174, 744–757.e24. [Google Scholar] [CrossRef] [PubMed]
  6. Quinodoz, S.A.; Bhat, P.; Chovanec, P.; Jachowicz, J.W.; 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. [Google Scholar] [CrossRef] [PubMed]
  7. Bickmore, W.A.; van Steensel, B. Genome Architecture: Domain Organization of Interphase Chromosomes. Cell 2013, 152, 1270–1284. [Google Scholar] [CrossRef]
  8. Dekker, J.; Mirny, L. The 3D Genome as Moderator of Chromosomal Communication. Cell 2016, 164, 1110–1121. [Google Scholar] [CrossRef]
  9. Dixon, J.R.; Gorkin, D.U.; Ren, B. Chromatin Domains: The Unit of Chromosome Organization. Mol. Cell 2016, 62, 668–680. [Google Scholar] [CrossRef]
  10. Spielmann, M.; Lupiáñez, D.G.; Mundlos, S. Structural Variation in the 3D Genome. Nat. Rev. Genet. 2018, 19, 453–467. [Google Scholar] [CrossRef]
  11. Akgol Oksuz, B.; Yang, L.; Abraham, S.; Venev, S.V.; Krietenstein, N.; Parsi, K.M.; Ozadam, H.; Oomen, M.E.; Nand, A.; Mao, H.; et al. Systematic Evaluation of Chromosome Conformation Capture Assays. Nat. Methods 2021, 18, 1046–1055. [Google Scholar] [CrossRef] [PubMed]
  12. Bickmore, W.A. The Spatial Organization of the Human Genome. Annu. Rev. Genom. Hum. Genet. 2013, 14, 67–84. [Google Scholar] [CrossRef] [PubMed]
  13. Bonev, B.; Mendelson Cohen, N.; Szabo, Q.; Fritsch, L.; Papadopoulos, G.L.; Lubling, Y.; Xu, X.; Lv, X.; Hugnot, J.-P.; Tanay, A.; et al. Multiscale 3D Genome Rewiring during Mouse Neural Development. Cell 2017, 171, 557–572.e24. [Google Scholar] [CrossRef] [PubMed]
  14. Rao, S.S.P.; Huntley, M.H.; Durand, N.C.; Stamenova, E.K.; Bochkov, I.D.; Robinson, J.T.; Sanborn, A.L.; Machol, I.; Omer, A.D.; Lander, E.S.; et al. A 3D Map of the Human Genome at Kilobase Resolution Reveals Principles of Chromatin Looping. Cell 2014, 159, 1665–1680. [Google Scholar] [CrossRef] [PubMed]
  15. Nora, E.P.; Lajoie, B.R.; Schulz, E.G.; Giorgetti, L.; Okamoto, I.; Servant, N.; Piolot, T.; van Berkum, N.L.; Meisig, J.; Sedat, J.; et al. Spatial Partitioning of the Regulatory Landscape of the X-Inactivation Centre. Nature 2012, 485, 381–385. [Google Scholar] [CrossRef]
  16. Dixon, J.R.; Selvaraj, S.; Yue, F.; Kim, A.; Li, Y.; Shen, Y.; Hu, M.; Liu, J.S.; Ren, B. Topological Domains in Mammalian Genomes Identified by Analysis of Chromatin Interactions. Nature 2012, 485, 376–380. [Google Scholar] [CrossRef]
  17. Fraser, J.; Ferrai, C.; Chiariello, A.M.; Schueler, M.; Rito, T.; Laudanno, G.; Barbieri, M.; Moore, B.L.; Kraemer, D.C.; Aitken, S.; et al. Hierarchical Folding and Reorganization of Chromosomes Are Linked to Transcriptional Changes in Cellular Differentiation. Mol. Syst. Biol. 2015, 11, 852. [Google Scholar] [CrossRef]
  18. van Steensel, B.; Belmont, A.S. Lamina-Associated Domains: Links with Chromosome Architecture, Heterochromatin, and Gene Repression. Cell 2017, 169, 780–791. [Google Scholar] [CrossRef]
  19. Cremer, T.; Cremer, C. Chromosome Territories, Nuclear Architecture and Gene Regulation in Mammalian Cells. Nat. Rev. Genet. 2001, 2, 292–301. [Google Scholar] [CrossRef]
  20. Sexton, T.; Cavalli, G. The Role of Chromosome Domains in Shaping the Functional Genome. Cell 2015, 160, 1049–1059. [Google Scholar] [CrossRef]
  21. Schoenfelder, S.; Fraser, P. Long-Range Enhancer–Promoter Contacts in Gene Expression Control. Nat. Rev. Genet. 2019, 20, 437–455. [Google Scholar] [CrossRef] [PubMed]
  22. Dowen, J.M.; Fan, Z.P.; Hnisz, D.; Ren, G.; Abraham, B.J.; Zhang, L.N.; Weintraub, A.S.; Schuijers, J.; Lee, T.I.; Zhao, K.; et al. Control of Cell Identity Genes Occurs in Insulated Neighborhoods in Mammalian Chromosomes. Cell 2014, 159, 374–387. [Google Scholar] [CrossRef] [PubMed]
  23. Bonev, B.; Cavalli, G. Organization and Function of the 3D Genome. Nat. Rev. Genet. 2016, 17, 661–678. [Google Scholar] [CrossRef] [PubMed]
  24. Kubo, N.; Ishii, H.; Xiong, X.; Bianco, S.; Meitinger, F.; Hu, R.; Hocker, J.D.; Conte, M.; Gorkin, D.; Yu, M.; et al. Promoter-Proximal CTCF Binding Promotes Distal Enhancer-Dependent Gene Activation. Nat. Struct. Mol. Biol. 2021, 28, 152–161. [Google Scholar] [CrossRef] [PubMed]
  25. Huang, H.; Zhu, Q.; Jussila, A.; Han, Y.; Bintu, B.; Kern, C.; Conte, M.; Zhang, Y.; Bianco, S.; Chiariello, A.M.; et al. CTCF Mediates Dosage- and Sequence-Context-Dependent Transcriptional Insulation by Forming Local Chromatin Domains. Nat. Genet. 2021, 53, 1064–1074. [Google Scholar] [CrossRef] [PubMed]
  26. Willemin, A.; Szabó, D.; Pombo, A. Epigenetic Regulatory Layers in the 3D Nucleus. Mol. Cell 2024, 84, 415–428. [Google Scholar] [CrossRef] [PubMed]
  27. Salamon, I.; Serio, S.; Bianco, S.; Pagiatakis, C.; Crasto, S.; Chiariello, A.M.; Conte, M.; Cattaneo, P.; Fiorillo, L.; Felicetta, A.; et al. Divergent Transcription of the Nkx2-5 Locus Generates Two Enhancer RNAs with Opposing Functions. iScience 2020, 23, 101539. [Google Scholar] [CrossRef]
  28. Misteli, T. The Self-Organizing Genome: Principles of Genome Architecture and Function. Cell 2020, 183, 28–45. [Google Scholar] [CrossRef]
  29. Gorkin, D.U.; Leung, D.; Ren, B. The 3D Genome in Transcriptional Regulation and Pluripotency. Cell Stem Cell 2014, 14, 762–775. [Google Scholar] [CrossRef]
  30. Cavalli, G.; Misteli, T. Functional Implications of Genome Topology. Nat. Struct. Mol. Biol. 2013, 20, 290–299. [Google Scholar] [CrossRef]
  31. Symmons, O.; Uslu, V.V.; Tsujimura, T.; Ruf, S.; Nassari, S.; Schwarzer, W.; Ettwiller, L.; Spitz, F. Functional and Topological Characteristics of Mammalian Regulatory Domains. Genome Res. 2014, 24, 390–400. [Google Scholar] [CrossRef] [PubMed]
  32. Lupiáñez, D.G.; Kraft, K.; Heinrich, V.; Krawitz, P.; Brancati, F.; Klopocki, E.; Horn, D.; Kayserili, H.; Opitz, J.M.; Laxova, R.; et al. Disruptions of Topological Chromatin Domains Cause Pathogenic Rewiring of Gene-Enhancer Interactions. Cell 2015, 161, 1012–1025. [Google Scholar] [CrossRef] [PubMed]
  33. Lupiáñez, D.G.; Spielmann, M.; Mundlos, S. Breaking TADs: How Alterations of Chromatin Domains Result in Disease. Trends Genet. 2016, 32, 225–237. [Google Scholar] [CrossRef] [PubMed]
  34. Weischenfeldt, J.; Dubash, T.; Drainas, A.P.; Mardin, B.R.; Chen, Y.; Stütz, A.M.; Waszak, S.M.; Bosco, G.; Halvorsen, A.R.; Raeder, B.; et al. Pan-Cancer Analysis of Somatic Copy-Number Alterations Implicates IRS4 and IGF2 in Enhancer Hijacking. Nat. Genet. 2017, 49, 65–74. [Google Scholar] [CrossRef] [PubMed]
  35. Kraft, K.; Magg, A.; Heinrich, V.; Riemenschneider, C.; Schöpflin, R.; Markowski, J.; Ibrahim, D.M.; Acuna-Hidalgo, R.; Despang, A.; Andrey, G.; et al. Serial Genomic Inversions Induce Tissue-Specific Architectural Stripes, Gene Misexpression and Congenital Malformations. Nat. Cell Biol. 2019, 21, 305–310. [Google Scholar] [CrossRef] [PubMed]
  36. Hnisz, D.; Weintraub, A.S.; Day, D.S.; Valton, A.-L.; Bak, R.O.; Li, C.H.; Goldmann, J.; Lajoie, B.R.; Fan, Z.P.; Sigova, A.A.; et al. Activation of Proto-Oncogenes by Disruption of Chromosome Neighborhoods. Science 2016, 351, 1454–1458. [Google Scholar] [CrossRef]
  37. Franke, M.; Ibrahim, D.M.; Andrey, G.; Schwarzer, W.; Heinrich, V.; Schöpflin, R.; Kraft, K.; Kempfer, R.; Jerković, I.; Chan, W.-L.; et al. Formation of New Chromatin Domains Determines Pathogenicity of Genomic Duplications. Nature 2016, 538, 265–269. [Google Scholar] [CrossRef]
  38. Sati, S.; Bonev, B.; Szabo, Q.; Jost, D.; Bensadoun, P.; Serra, F.; Loubiere, V.; Papadopoulos, G.L.; Rivera-Mulia, J.-C.; Fritsch, L.; et al. 4D Genome Rewiring during Oncogene-Induced and Replicative Senescence. Mol. Cell 2020, 78, 522–538.e9. [Google Scholar] [CrossRef]
  39. Anania, C.; Lupiáñez, D.G. Order and Disorder: Abnormal 3D Chromatin Organization in Human Disease. Brief. Funct. Genom. 2020, 19, 128–138. [Google Scholar] [CrossRef]
  40. Norton, H.K.; Phillips-Cremins, J.E. Crossed Wires: 3D Genome Misfolding in Human Disease. J. Cell Biol. 2017, 216, 3441–3452. [Google Scholar] [CrossRef]
  41. Ibn-Salem, J.; Köhler, S.; Love, M.I.; Chung, H.-R.; Huang, N.; Hurles, M.E.; Haendel, M.; Washington, N.L.; Smedley, D.; Mungall, C.J.; et al. Deletions of Chromosomal Regulatory Boundaries Are Associated with Congenital Disease. Genome Biol. 2014, 15, 423. [Google Scholar] [CrossRef] [PubMed]
  42. Flavahan, W.A.; Drier, Y.; Liau, B.B.; Gillespie, S.M.; Venteicher, A.S.; Stemmer-Rachamimov, A.O.; Suvà, M.L.; Bernstein, B.E. Insulator Dysfunction and Oncogene Activation in IDH Mutant Gliomas. Nature 2016, 529, 110–114. [Google Scholar] [CrossRef] [PubMed]
  43. Northcott, P.A.; Lee, C.; Zichner, T.; Stütz, A.M.; Erkek, S.; Kawauchi, D.; Shih, D.J.H.; Hovestadt, V.; Zapatka, M.; Sturm, D.; et al. Enhancer Hijacking Activates GFI1 Family Oncogenes in Medulloblastoma. Nature 2014, 511, 428–434. [Google Scholar] [CrossRef] [PubMed]
  44. Finn, E.H.; Misteli, T. Molecular Basis and Biological Function of Variability in Spatial Genome Organization. Science 2019, 365, eaaw9498. [Google Scholar] [CrossRef]
  45. Bintu, B.; Mateo, L.J.; Su, J.-H.; Sinnott-Armstrong, N.A.; Parker, M.; Kinrot, S.; Yamaya, K.; Boettiger, A.N.; Zhuang, X. Super-Resolution Chromatin Tracing Reveals Domains and Cooperative Interactions in Single Cells. Science 2018, 362, eaau1783. [Google Scholar] [CrossRef]
  46. Su, J.-H.; Zheng, P.; Kinrot, S.S.; Bintu, B.; Zhuang, X. Genome-Scale Imaging of the 3D Organization and Transcriptional Activity of Chromatin. Cell 2020, 182, 1641–1659.e26. [Google Scholar] [CrossRef]
  47. Mateo, L.J.; Murphy, S.E.; Hafner, A.; Cinquini, I.S.; Walker, C.A.; Boettiger, A.N. Visualizing DNA Folding and RNA in Embryos at Single-Cell Resolution. Nature 2019, 568, 49–54. [Google Scholar] [CrossRef]
  48. Wang, S.; Su, J.H.; Beliveau, B.J.; Bintu, B.; Moffitt, J.R.; Wu, C.T.; Zhuang, X. Spatial Organization of Chromatin Domains and Compartments in Single Chromosomes. Science 2016, 353. [Google Scholar] [CrossRef]
  49. Nguyen, H.Q.; Chattoraj, S.; Castillo, D.; Nguyen, S.C.; Nir, G.; Lioutas, A.; Hershberg, E.A.; Martins, N.M.C.; Reginato, P.L.; Hannan, M.; et al. 3D Mapping and Accelerated Super-Resolution Imaging of the Human Genome Using in Situ Sequencing. Nat. Methods 2020, 17, 822–832. [Google Scholar] [CrossRef]
  50. Takei, Y.; Yun, J.; Zheng, S.; Ollikainen, N.; Pierson, N.; White, J.; Shah, S.; Thomassie, J.; Suo, S.; Eng, C.-H.L.; et al. Integrated Spatial Genomics Reveals Global Architecture of Single Nuclei. Nature 2021, 590, 344–350. [Google Scholar] [CrossRef]
  51. Takei, Y.; Zheng, S.; Yun, J.; Shah, S.; Pierson, N.; White, J.; Schindler, S.; Tischbirek, C.H.; Yuan, G.-C.; Cai, L. Single-Cell Nuclear Architecture across Cell Types in the Mouse Brain. Science 2021, 374, 586–594. [Google Scholar] [CrossRef] [PubMed]
  52. Szabo, Q.; Donjon, A.; Jerković, I.; Papadopoulos, G.L.; Cheutin, T.; Bonev, B.; Nora, E.P.; Bruneau, B.G.; Bantignies, F.; Cavalli, G. Regulation of Single-Cell Genome Organization into TADs and Chromatin Nanodomains. Nat. Genet. 2020, 52, 1151–1157. [Google Scholar] [CrossRef] [PubMed]
  53. Finn, E.H.; Pegoraro, G.; Brandão, H.B.; Valton, A.-L.; Oomen, M.E.; Dekker, J.; Mirny, L.; Misteli, T. Extensive Heterogeneity and Intrinsic Variation in Spatial Genome Organization. Cell 2019, 176, 1502–1515.e10. [Google Scholar] [CrossRef] [PubMed]
  54. Boettiger, A.N.; Bintu, B.; Moffitt, J.R.; Wang, S.; Beliveau, B.J.; Fudenberg, G.; Imakaev, M.; Mirny, L.A.; Wu, C.; Zhuang, X. Super-Resolution Imaging Reveals Distinct Chromatin Folding for Different Epigenetic States. Nature 2016, 529, 418–422. [Google Scholar] [CrossRef] [PubMed]
  55. Götz, M.; Messina, O.; Espinola, S.; Fiche, J.-B.; Nollmann, M. Multiple Parameters Shape the 3D Chromatin Structure of Single Nuclei at the Doc Locus in Drosophila. Nat. Commun. 2022, 13, 5375. [Google Scholar] [CrossRef] [PubMed]
  56. Cardozo Gizzi, A.M.; Cattoni, D.I.; Fiche, J.-B.; Espinola, S.M.; Gurgo, J.; Messina, O.; Houbron, C.; Ogiyama, Y.; Papadopoulos, G.L.; 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.e5. [Google Scholar] [CrossRef]
  57. Nir, G.; Farabella, I.; Pérez Estrada, C.; Ebeling, C.G.; Beliveau, B.J.; Sasaki, H.M.; Lee, S.H.; Nguyen, S.C.; McCole, R.B.; Chattoraj, S.; et al. Walking along Chromosomes with Super-Resolution Imaging, Contact Maps, and Integrative Modeling. PLoS Genet. 2018, 14, e1007872. [Google Scholar] [CrossRef]
  58. Jerkovic’, I.; Cavalli, G. Understanding 3D Genome Organization by Multidisciplinary Methods. Nat. Rev. Mol. Cell Biol. 2021, 22, 511–528. [Google Scholar] [CrossRef]
  59. Li, Q.; Tjong, H.; Li, X.; Gong, K.; Zhou, X.J.; Chiolo, I.; Alber, F. The Three-Dimensional Genome Organization of Drosophila Melanogaster through Data Integration. Genome Biol. 2017, 18, 145. [Google Scholar] [CrossRef]
  60. Tjong, H.; Li, W.; Kalhor, R.; Dai, C.; Hao, S.; Gong, K.; Zhou, Y.; Li, H.; Zhou, X.J.; Le Gros, M.A.; et al. Population-Based 3D Genome Structure Analysis Reveals Driving Forces in Spatial Genome Organization. Proc. Natl. Acad. Sci. USA 2016, 113, E1663–E1672. [Google Scholar] [CrossRef]
  61. Chiariello, A.M.; Annunziatella, C.; Bianco, S.; Esposito, A.; Nicodemi, M. Polymer Physics of Chromosome Large-Scale 3D Organisation. Sci. Rep. 2016, 6, 29775. [Google Scholar] [CrossRef] [PubMed]
  62. Chiariello, A.M.; Bianco, S.; Esposito, A.; Fiorillo, L.; Conte, M.; Irani, E.; Musella, F.; Abraham, A.; Prisco, A.; Nicodemi, M. Physical Mechanisms of Chromatin Spatial Organization. FEBS J. 2022, 289, 1180–1190. [Google Scholar] [CrossRef] [PubMed]
  63. Brackley, C.A.; Taylor, S.; Papantonis, A.; Cook, P.R.; Marenduzzo, D. Nonspecific Bridging-Induced Attraction Drives Clustering of DNA-Binding Proteins and Genome Organization. Proc. Natl. Acad. Sci. USA 2013, 110, E3605–E3611. [Google Scholar] [CrossRef] [PubMed]
  64. Brackley, C.A.; Liebchen, B.; Michieletto, D.; Mouvet, F.; Cook, P.R.; Marenduzzo, D. Ephemeral Protein Binding to DNA Shapes Stable Nuclear Bodies and Chromatin Domains. Biophys. J. 2017, 112, 1085–1093. [Google Scholar] [CrossRef] [PubMed]
  65. Di Pierro, M.; Zhang, B.; Aiden, E.L.; Wolynes, P.G.; Onuchic, J.N. Transferable Model for Chromosome Architecture. Proc. Natl. Acad. Sci. USA 2016, 113, 12168–12173. [Google Scholar] [CrossRef]
  66. Rosa, A.; Everaers, R. Structure and Dynamics of Interphase Chromosomes. PLoS Comput. Biol. 2008, 4, e1000153. [Google Scholar] [CrossRef]
  67. Esposito, A.; Bianco, S.; Fiorillo, L.; Conte, M.; Abraham, A.; Musella, F.; Nicodemi, M.; Prisco, A.; Chiariello, A.M. Polymer Models Are a Versatile Tool to Study Chromatin 3D Organization. Biochem. Soc. Trans. 2021, 49, 1675–1684. [Google Scholar] [CrossRef]
  68. Bianco, S.; Chiariello, A.M.; Conte, M.; Esposito, A.; Fiorillo, L.; Musella, F.; Nicodemi, M. Computational Approaches from Polymer Physics to Investigate Chromatin Folding. Curr. Opin. Cell Biol. 2020, 64, 10–17. [Google Scholar] [CrossRef]
  69. Neguembor, M.V.; Arcon, J.P.; Buitrago, D.; Lema, R.; Walther, J.; Garate, X.; Martin, L.; Romero, P.; AlHaj Abed, J.; Gut, M.; et al. MiOS, an Integrated Imaging and Computational Strategy to Model Gene Folding with Nucleosome Resolution. Nat. Struct. Mol. Biol. 2022, 29, 1011–1023. [Google Scholar] [CrossRef]
  70. Racko, D.; Benedetti, F.; Dorier, J.; Stasiak, A. Transcription-Induced Supercoiling as the Driving Force of Chromatin Loop Extrusion during Formation of TADs in Interphase Chromosomes. Nucleic Acids Res. 2018, 46, 1648–1660. [Google Scholar] [CrossRef]
  71. Serra, F.; Baù, D.; Goodstadt, M.; Castillo, D.; Filion, G.J.; Marti-Renom, M.A. Automatic Analysis and 3D-Modelling of Hi-C Data Using TADbit Reveals Structural Features of the Fly Chromatin Colors. PLoS Comput. Biol. 2017, 13, e1005665. [Google Scholar] [CrossRef] [PubMed]
  72. Sanborn, A.L.; Rao, S.S.P.; Huang, S.C.; Durand, N.C.; Huntley, M.H.; Jewett, A.I.; Bochkov, I.D.; Chinnappan, D.; Cutkosky, A.; Li, J.; et al. Chromatin Extrusion Explains Key Features of Loop and Domain Formation in Wild-Type and Engineered Genomes. Proc. Natl. Acad. Sci. USA 2015, 112, E6456–E6465. [Google Scholar] [CrossRef] [PubMed]
  73. Fudenberg, G.; Imakaev, M.; Lu, C.; Goloborodko, A.; Abdennur, N.; Mirny, L.A. Formation of Chromosomal Domains by Loop Extrusion. Cell Rep. 2016, 15, 2038–2049. [Google Scholar] [CrossRef] [PubMed]
  74. Yildirim, A.; Boninsegna, L.; Zhan, Y.; Alber, F. Uncovering the Principles of Genome Folding by 3D Chromatin Modeling. Cold Spring Harb. Perspect. Biol. 2022, 14, a039693. [Google Scholar] [CrossRef] [PubMed]
  75. Boninsegna, L.; Yildirim, A.; Polles, G.; Zhan, Y.; Quinodoz, S.A.; Finn, E.H.; Guttman, M.; Zhou, X.J.; Alber, F. Integrative Genome Modeling Platform Reveals Essentiality of Rare Contact Events in 3D Genome Organizations. Nat. Methods 2022, 19, 938–949. [Google Scholar] [CrossRef]
  76. Esposito, A.; Abraham, A.; Conte, M.; Vercellone, F.; Prisco, A.; Bianco, S.; Chiariello, A.M. The Physics of DNA Folding: Polymer Models and Phase-Separation. Polymers 2022, 14, 1918. [Google Scholar] [CrossRef]
  77. Conte, M.; Irani, E.; Chiariello, A.M.; Abraham, A.; Bianco, S.; Esposito, A.; Nicodemi, M. Loop-Extrusion and Polymer Phase-Separation Can Co-Exist at the Single-Molecule Level to Shape Chromatin Folding. Nat. Commun. 2022, 13, 4070. [Google Scholar] [CrossRef]
  78. Fiorillo, L.; Bianco, S.; Esposito, A.; Conte, M.; Sciarretta, R.; Musella, F.; Chiariello, A.M. A Modern Challenge of Polymer Physics: Novel Ways to Study, Interpret, and Reconstruct Chromatin Structure. WIREs Comput. Mol. Sci. 2020, 10, e1454. [Google Scholar] [CrossRef]
  79. Esposito, A.; Chiariello, A.M.; Conte, M.; Fiorillo, L.; Musella, F.; Sciarretta, R.; Bianco, S. Higher-Order Chromosome Structures Investigated by Polymer Physics in Cellular Morphogenesis and Differentiation. J. Mol. Biol. 2020, 432, 701–711. [Google Scholar] [CrossRef]
  80. Zhang, B.; Wolynes, P.G. Topology, Structures, and Energy Landscapes of Human Chromosomes. Proc. Natl. Acad. Sci. USA 2015, 112, 6062–6067. [Google Scholar] [CrossRef]
  81. Shi, G.; Thirumalai, D. From Hi-C Contact Map to Three-Dimensional Organization of Interphase Human Chromosomes. Phys. Rev. X 2021, 11, 011051. [Google Scholar] [CrossRef]
  82. Lin, D.; Bonora, G.; Yardimci, G.G.; Noble, W.S. Computational Methods for Analyzing and Modeling Genome Structure and Organization. Wiley Interdiscip. Rev. Syst. Biol. Med. 2018, 11, e1435. [Google Scholar] [CrossRef] [PubMed]
  83. Shi, G.; Liu, L.; Hyeon, C.; Thirumalai, D. Interphase Human Chromosome Exhibits out of Equilibrium Glassy Dynamics. Nat. Commun. 2018, 9, 3161. [Google Scholar] [CrossRef] [PubMed]
  84. Plewczynski, D.; Kadlof, M. Computational Modelling of Three-Dimensional Genome Structure. Methods 2020, 181–182, 1–4. [Google Scholar] [CrossRef] [PubMed]
  85. Banigan, E.J.; Mirny, L.A. Loop Extrusion: Theory Meets Single-Molecule Experiments. Curr. Opin. Cell Biol. 2020, 64, 124–138. [Google Scholar] [CrossRef]
  86. Nicodemi, M.; Prisco, A. Thermodynamic Pathways to Genome Spatial Organization in the Cell Nucleus. Biophys. J. 2009, 96, 2168–2177. [Google Scholar] [CrossRef]
  87. Barbieri, M.; Chotalia, M.; Fraser, J.; Lavitas, L.-M.; Dostie, J.; Pombo, A.; Nicodemi, M. Complexity of Chromatin Folding Is Captured by the Strings and Binders Switch Model. Proc. Natl. Acad. Sci. USA 2012, 109, 16173–16178. [Google Scholar] [CrossRef]
  88. Crippa, M.; Zhan, Y.; Tiana, G. Effective Model of Loop Extrusion Predicts Chromosomal Domains. Phys. Rev. E 2020, 102, 032414. [Google Scholar] [CrossRef]
  89. Conte, M.; Esposito, A.; Vercellone, F.; Abraham, A.; Bianco, S. Unveiling the Machinery behind Chromosome Folding by Polymer Physics Modeling. Int. J. Mol. Sci. 2023, 24, 3660. [Google Scholar] [CrossRef]
  90. Buckle, A.; Brackley, C.A.; Boyle, S.; Marenduzzo, D.; Gilbert, N. Polymer Simulations of Heteromorphic Chromatin Predict the 3D Folding of Complex Genomic Loci. Mol. Cell 2018, 72, 786–797.e11. [Google Scholar] [CrossRef]
  91. Brackley, C.A.; Johnson, J.; Michieletto, D.; Morozov, A.N.; Nicodemi, M.; Cook, P.R.; Marenduzzo, D. Extrusion without a Motor: A New Take on the Loop Extrusion Model of Genome Organization. Nucleus 2018, 9, 95–103. [Google Scholar] [CrossRef] [PubMed]
  92. Conte, M.; Esposito, A.; Fiorillo, L.; Annunziatella, C.; Corrado, A.; Musella, F.; Sciarretta, R.; Chiariello, A.M.; Bianco, S. Hybrid Machine Learning and Polymer Physics Approach to Investigate 3D Chromatin Structure. In Lecture Notes in Artificial Intelligence and Lecture Notes in Bioinformatics, Proceedings of theEuro-Par 2019: Parallel Processing Workshops: Euro-Par 2019 International Workshops, Göttingen, Germany, 26–30 August 2019; Lecture Notes in Computer Science; Springer: Berlin/Heidelberg, Germany, 2020; Volume 11997. [Google Scholar]
  93. Di Stefano, M.; Paulsen, J.; Lien, T.G.; Hovig, E.; Micheletti, C. Hi-C-Constrained Physical Models of Human Chromosomes Recover Functionally-Related Properties of Genome Organization. Sci. Rep. 2016, 6, 35985. [Google Scholar] [CrossRef] [PubMed]
  94. Erdel, F.; Rippe, K. Formation of Chromatin Subcompartments by Phase Separation. Biophys. J. 2018, 114, 2262–2270. [Google Scholar] [CrossRef] [PubMed]
  95. Brackley, C.A.; Johnson, J.; Kelly, S.; Cook, P.R.; Marenduzzo, D. Simulated Binding of Transcription Factors to Active and Inactive Regions Folds Human Chromosomes into Loops, Rosettes and Topological Domains. Nucleic Acids Res. 2016, 44, 3503–3512. [Google Scholar] [CrossRef] [PubMed]
  96. Di Stefano, M.; Rosa, A.; Belcastro, V.; di Bernardo, D.; Micheletti, C. Colocalization of Coregulated Genes: A Steered Molecular Dynamics Study of Human Chromosome 19. PLoS Comput. Biol. 2013, 9, e1003019. [Google Scholar] [CrossRef]
  97. Di Stefano, M.; Stadhouders, R.; Farabella, I.; Castillo, D.; Serra, F.; Graf, T.; Marti-Renom, M.A. Transcriptional Activation during Cell Reprogramming Correlates with the Formation of 3D Open Chromatin Hubs. Nat. Commun. 2020, 11, 2564. [Google Scholar] [CrossRef]
  98. Lesne, A.; Riposo, J.; Roger, P.; Cournac, A.; Mozziconacci, J. 3D Genome Reconstruction from Chromosomal Contacts. Nat. Methods 2014, 11, 1141–1143. [Google Scholar] [CrossRef]
  99. Brackey, C.A.; Marenduzzo, D.; Gilbert, N. Mechanistic Modeling of Chromatin Folding to Understand Function. Nat. Methods 2020, 17, 767–775. [Google Scholar] [CrossRef]
  100. Zhang, S.; Chasman, D.; Knaack, S.; Roy, S. In Silico Prediction of High-Resolution Hi-C Interaction Matrices. Nat. Commun. 2019, 10, 5449. [Google Scholar] [CrossRef]
  101. Ghosh, S.K.; Jost, D. How Epigenome Drives Chromatin Folding and Dynamics, Insights from Efficient Coarse-Grained Models of Chromosomes. PLoS Comput. Biol. 2018, 14, e1006159. [Google Scholar] [CrossRef]
  102. Brackley, C.A.; Brown, J.M.; Waithe, D.; Babbs, C.; Davies, J.; Hughes, J.R.; Buckle, V.J.; Marenduzzo, D. Predicting the Three-Dimensional Folding of Cis-Regulatory Regions in Mammalian Genomes Using Bioinformatic Data and Polymer Models. Genome Biol. 2016, 17, 59. [Google Scholar] [CrossRef] [PubMed]
  103. Di Pierro, M.; Cheng, R.R.; Lieberman Aiden, E.; Wolynes, P.G.; Onuchic, J.N. De Novo Prediction of Human Chromosome Structures: Epigenetic Marking Patterns Encode Genome Architecture. Proc. Natl. Acad. Sci. USA 2017, 114, 12126–12131. [Google Scholar] [CrossRef] [PubMed]
  104. Belokopytova, P.; Viesná, E.; Chiliński, M.; Qi, Y.; Salari, H.; Di Stefano, M.; Esposito, A.; Conte, M.; Chiariello, A.M.; Teif, V.B.; et al. 3DGenBench: A Web-Server to Benchmark Computational Models for 3D Genomics. Nucleic Acids Res. 2022, 50, W4–W12. [Google Scholar] [CrossRef]
  105. Chiang, M.; Michieletto, D.; Brackley, C.A.; Rattanavirotkul, N.; Mohammed, H.; Marenduzzo, D.; Chandra, T. Polymer Modeling Predicts Chromosome Reorganization in Senescence. Cell Rep. 2019, 28, 3212–3223.e6. [Google Scholar] [CrossRef] [PubMed]
  106. Lin, X.; Qi, Y.; Latham, A.P.; Zhang, B. Multiscale Modeling of Genome Organization with Maximum Entropy Optimization. J. Chem. Phys. 2021, 155, 010901. [Google Scholar] [CrossRef]
  107. Michieletto, D.; Orlandini, E.; Marenduzzo, D. Polymer Model with Epigenetic Recoloring Reveals a Pathway for the de Novo Establishment and 3D Organization of Chromatin Domains. Phys. Rev. X 2016, 6, 041047. [Google Scholar] [CrossRef]
  108. Laghmach, R.; Di Pierro, M.; Potoyan, D.A. Mesoscale Liquid Model of Chromatin Recapitulates Nuclear Order of Eukaryotes. Biophys. J. 2020, 118, 2130–2140. [Google Scholar] [CrossRef]
  109. Salari, H.; Di Stefano, M.; Jost, D. Spatial Organization of Chromosomes Leads to Heterogeneous Chromatin Motion and Drives the Liquid- or Gel-like Dynamical Behavior of Chromatin. Genome Res. 2022, 32, 28–43. [Google Scholar] [CrossRef]
  110. Jost, D.; Carrivain, P.; Cavalli, G.; Vaillant, C. Modeling Epigenome Folding: Formation and Dynamics of Topologically Associated Chromatin Domains. Nucleic Acids Res. 2014, 42, 9553–9561. [Google Scholar] [CrossRef]
  111. Tortora, M.M.; Salari, H.; Jost, D. Chromosome Dynamics during Interphase: A Biophysical Perspective. Curr. Opin. Genet. Dev. 2020, 61, 37–43. [Google Scholar] [CrossRef]
  112. Bohn, M.; Heermann, D.W. Diffusion-Driven Looping Provides a Consistent Framework for Chromatin Organization. PLoS ONE 2010, 5, e12218. [Google Scholar] [CrossRef] [PubMed]
  113. Conte, M.; Fiorillo, L.; Bianco, S.; Chiariello, A.M.; Esposito, A.; Musella, F.; Flora, F.; Abraham, A.; Nicodemi, M. A Polymer Physics Model to Dissect Genome Organization in Healthy and Pathological Phenotypes. In Methods in Molecular Biology; Springer: Berlin/Heidelberg, Germany, 2022; Volume 2301, pp. 307–316. [Google Scholar]
  114. Fudenberg, G.; Kelley, D.R.; Pollard, K.S. Predicting 3D Genome Folding from DNA Sequence with Akita. Nat. Methods 2020, 17, 1111–1117. [Google Scholar] [CrossRef] [PubMed]
  115. Zhou, J. Sequence-Based Modeling of Three-Dimensional Genome Architecture from Kilobase to Chromosome Scale. Nat. Genet. 2022, 54, 725–734. [Google Scholar] [CrossRef] [PubMed]
  116. Singh, S.; Yang, Y.; Póczos, B.; Ma, J. Predicting Enhancer-promoter Interaction from Genomic Sequence with Deep Neural Networks. Quant. Biol. 2019, 7, 122–137. [Google Scholar] [CrossRef]
  117. Bianco, S.; Lupiáñez, D.G.; Chiariello, A.M.; Annunziatella, C.; Kraft, K.; Schöpflin, R.; Wittler, L.; Andrey, G.; Vingron, M.; Pombo, A.; et al. Polymer Physics Predicts the Effects of Structural Variants on Chromatin Architecture. Nat. Genet. 2018, 50, 662–667. [Google Scholar] [CrossRef]
  118. Chiariello, A.M.; Abraham, A.; Bianco, S.; Esposito, A.; Fontana, A.; Vercellone, F.; Conte, M.; Nicodemi, M. Multiscale Modelling of Chromatin 4D Organization in SARS-CoV-2 Infected Cells. Nat. Commun. 2024, 15, 4014. [Google Scholar] [CrossRef]
  119. Schwessinger, R.; Gosden, M.; Downes, D.; Brown, R.C.; Oudelaar, A.M.; Telenius, J.; Teh, Y.W.; Lunter, G.; Hughes, J.R. DeepC: Predicting 3D Genome Folding Using Megabase-Scale Transfer Learning. Nat. Methods 2020, 17, 1118–1124. [Google Scholar] [CrossRef]
  120. Li, W.; Wong, W.H.; Jiang, R. DeepTACT: Predicting 3D Chromatin Contacts via Bootstrapping Deep Learning. Nucleic Acids Res. 2019, 47, e60. [Google Scholar] [CrossRef]
  121. Belokopytova, P.S.; Nuriddinov, M.A.; Mozheiko, E.A.; Fishman, D.; Fishman, V. Quantitative Prediction of Enhancer–Promoter Interactions. Genome Res. 2020, 30, 72–84. [Google Scholar] [CrossRef]
  122. Zhang, R.; Wang, Y.; Yang, Y.; Zhang, Y.; Ma, J. Predicting CTCF-Mediated Chromatin Loops Using CTCF-MP. Bioinformatics 2018, 34, i133–i141. [Google Scholar] [CrossRef]
  123. Tan, J.; Shenker-Tauris, N.; Rodriguez-Hernaez, J.; Wang, E.; Sakellaropoulos, T.; Boccalatte, F.; Thandapani, P.; Skok, J.; Aifantis, I.; Fenyö, D.; et al. Cell-Type-Specific Prediction of 3D Chromatin Organization Enables High-Throughput in Silico Genetic Screening. Nat. Biotechnol. 2023, 41, 1140–1150. [Google Scholar] [CrossRef] [PubMed]
  124. Cao, F.; Zhang, Y.; Cai, Y.; Animesh, S.; Zhang, Y.; Akincilar, S.C.; Loh, Y.P.; Li, X.; Chng, W.J.; Tergaonkar, V.; et al. Chromatin Interaction Neural Network (ChINN): A Machine Learning-Based Method for Predicting Chromatin Interactions from DNA Sequences. Genome Biol. 2021, 22, 226. [Google Scholar] [CrossRef] [PubMed]
  125. Fiorillo, L.; Musella, F.; Conte, M.; Kempfer, R.; Chiariello, A.M.; Bianco, S.; Kukalev, A.; Irastorza-Azcarate, I.; Esposito, A.; Abraham, A.; et al. Comparison of the Hi-C, GAM and SPRITE Methods Using Polymer Models of Chromatin. Nat. Methods 2021, 18, 482–490. [Google Scholar] [CrossRef] [PubMed]
  126. Fiorillo, L.; Bianco, S.; Chiariello, A.M.; Barbieri, M.; Esposito, A.; Annunziatella, C.; Conte, M.; Corrado, A.; Prisco, A.; Pombo, A.; et al. Inference of Chromosome 3D Structures from GAM Data by a Physics Computational Approach. Methods 2020, 181–182, 70–79. [Google Scholar] [CrossRef]
  127. Conte, M.; Fiorillo, L.; Bianco, S.; Chiariello, A.M.; Esposito, A.; Nicodemi, M. Polymer Physics Indicates Chromatin Folding Variability across Single-Cells Results from State Degeneracy in Phase Separation. Nat. Commun. 2020, 11, 3289. [Google Scholar] [CrossRef]
  128. Conte, M.; Chiariello, A.M.; Bianco, S.; Esposito, A.; Abraham, A.; Nicodemi, M. Physics-Based Polymer Models to Probe Chromosome Structure in Single Molecules. In Methods in Molecular Biology; Humana: New York, NY, USA; Springer: Berlin/Heidelberg, Germany, 2023; Volume 2655. [Google Scholar] [CrossRef]
  129. Defenouillère, Q.; Yao, Y.; Mouaikel, J.; Namane, A.; Galopier, A.; Decourty, L.; Doyen, A.; Malabat, C.; Saveanu, C.; Jacquier, A.; et al. Cdc48-Associated Complex Bound to 60S Particles Is Required for the Clearance of Aberrant Translation Products. Proc. Natl. Acad. Sci. USA 2013, 110, 5046–5051. [Google Scholar] [CrossRef]
  130. Joazeiro, C.A.P. Mechanisms and Functions of Ribosome-Associated Protein Quality Control. Nat. Rev. Mol. Cell Biol. 2019, 20, 368–383. [Google Scholar] [CrossRef]
  131. Chu, J.; Hong, N.A.; Masuda, C.A.; Jenkins, B.V.; Nelms, K.A.; Goodnow, C.C.; Glynne, R.J.; Wu, H.; Masliah, E.; Joazeiro, C.A.P.; et al. A Mouse Forward Genetics Screen Identifies LISTERIN as an E3 Ubiquitin Ligase Involved in Neurodegeneration. Proc. Natl. Acad. Sci. USA 2009, 106, 2097–2103. [Google Scholar] [CrossRef]
  132. Ghosh, A.; Shcherbik, N. Cooperativity between the Ribosome-Associated Chaperone Ssb/RAC and the Ubiquitin Ligase Ltn1 in Ubiquitination of Nascent Polypeptides. Int. J. Mol. Sci. 2020, 21, 6815. [Google Scholar] [CrossRef]
  133. Farhang, S.; Sabaie, H.; Gharesouran, J.; Asadi, M.R.; Arsang-Jang, S.; Ghafouri-Fard, S.; Taheri, M.; Rezazadeh, M. Expression Analysis of Ermin and Listerin E3 Ubiquitin Protein Ligase 1 Genes in the Periphery of Patients with Schizophrenia. J. Mol. Neurosci. 2022, 72, 246–254. [Google Scholar] [CrossRef]
  134. Bengtson, M.H.; Joazeiro, C.A.P. Role of a Ribosome-Associated E3 Ubiquitin Ligase in Protein Quality Control. Nature 2010, 467, 470–473. [Google Scholar] [CrossRef] [PubMed]
  135. Shao, S.; Brown, A.; Santhanam, B.; Hegde, R.S. Structure and Assembly Pathway of the Ribosome Quality Control Complex. Mol. Cell 2015, 57, 433–444. [Google Scholar] [CrossRef] [PubMed]
  136. Matsuo, Y.; Ikeuchi, K.; Saeki, Y.; Iwasaki, S.; Schmidt, C.; Udagawa, T.; Sato, F.; Tsuchiya, H.; Becker, T.; Tanaka, K.; et al. Ubiquitination of Stalled Ribosome Triggers Ribosome-Associated Quality Control. Nat. Commun. 2017, 8, 159. [Google Scholar] [CrossRef] [PubMed]
  137. Belaghzal, H.; Dekker, J.; Gibcus, J.H. Hi-C 2.0: An Optimized Hi-C Procedure for High-Resolution Genome-Wide Mapping of Chromosome Conformation. Methods 2017, 123, 56–65. [Google Scholar] [CrossRef] [PubMed]
  138. Reiff, S.B.; Schroeder, A.J.; Kırlı, K.; Cosolo, A.; Bakker, C.; Lee, S.; Veit, A.D.; Balashov, A.K.; Vitzthum, C.; Ronchetti, W.; et al. The 4D Nucleome Data Portal as a Resource for Searching and Visualizing Curated Nucleomics Data. Nat. Commun. 2022, 13, 2365. [Google Scholar] [CrossRef]
  139. Dekker, J.; Belmont, A.S.; Guttman, M.; Leshyk, V.O.; Lis, J.T.; Lomvardas, S.; Mirny, L.A.; O’Shea, C.C.; Park, P.J.; Ren, B.; et al. The 4D Nucleome Project. Nature 2017, 549, 219–226. [Google Scholar] [CrossRef]
  140. De Gennes, P.G. Scaling Concepts in Polymer Physics; Cornell University Press: Ithaca, NY, USA, 1979; ISBN 080141203X. [Google Scholar]
  141. Shrinivas, K.; Sabari, B.R.; Coffey, E.L.; Klein, I.A.; Boija, A.; Zamudio, A.V.; Schuijers, J.; Hannett, N.M.; Sharp, P.A.; Young, R.A.; et al. Enhancer Features That Drive Formation of Transcriptional Condensates. Mol. Cell 2019, 75, 549–561.e7. [Google Scholar] [CrossRef]
  142. Tafuri, F.; Chiariello, A.M. The Effect of Configurational Complexity in Hetero-Polymers on the Coil-Globule Phase Transition. Eur. Phys. J. Plus 2023, 138, 150. [Google Scholar] [CrossRef]
  143. Esposito, A.; Bianco, S.; Chiariello, A.M.; Abraham, A.; Fiorillo, L.; Conte, M.; Campanile, R.; Nicodemi, M. Polymer Physics Reveals a Combinatorial Code Linking 3D Chromatin Architecture to 1D Chromatin States. Cell Rep. 2022, 38, 110601. [Google Scholar] [CrossRef]
  144. Thompson, A.P.; Aktulga, H.M.; Berger, R.; Bolintineanu, D.S.; Brown, W.M.; Crozier, P.S.; in ’t Veld, P.J.; Kohlmeyer, A.; Moore, S.G.; Nguyen, T.D.; et al. LAMMPS—A Flexible Simulation Tool for Particle-Based Materials Modeling at the Atomic, Meso, and Continuum Scales. Comput. Phys. Commun. 2022, 271, 108171. [Google Scholar] [CrossRef]
  145. Plimpton, S. Fast Parallel Algorithms for Short-Range Molecular Dynamics. J. Comput. Phys. 1995, 117, 1–19. [Google Scholar] [CrossRef]
  146. Conte, M.; Esposito, A.; Fiorillo, L.; Campanile, R.; Annunziatella, C.; Corrado, A.; Chiariello, M.G.; Bianco, S.; Chiariello, A.M. Efficient Computational Implementation of Polymer Physics Models to Explore Chromatin Structure. Int. J. Parallel Emergent Distrib. Syst. 2022, 37, 91–102. [Google Scholar] [CrossRef]
  147. Kremer, K.; Grest, G.S. Dynamics of Entangled Linear Polymer Melts: A Molecular-Dynamics Simulation. J. Chem. Phys. 1990, 92, 5057–5086. [Google Scholar] [CrossRef]
  148. Allen, M.P.; Tildesley, D.J. Computer Simulation of Liquids (Oxford Science Publications); Oxford University Press: Oxford, UK, 1989. [Google Scholar]
  149. Conte, M.; Fiorillo, L.; Annunziatella, C.; Esposito, A.; Musella, F.; Abraham, A.; Bianco, S.; Chiariello, A.M. Dynamic and Equilibrium Properties of Finite-Size Polymer Models of Chromosome Folding. Phys. Rev. E 2021, 104, 054402. [Google Scholar] [CrossRef] [PubMed]
  150. Chiariello, A.M.; Corberi, F.; Salerno, M. The Interplay between Phase Separation and Gene-Enhancer Communication: A Theoretical Study. Biophys. J. 2020, 119, 873–883. [Google Scholar] [CrossRef] [PubMed]
  151. Baum, M.; Erdel, F.; Wachsmuth, M.; Rippe, K. Retrieving the Intracellular Topology from Multi-Scale Protein Mobility Mapping in Living Cells. Nat. Commun. 2014, 5, 4494. [Google Scholar] [CrossRef]
  152. Conte, M.; Chiariello, A.M.; Abraham, A.; Bianco, S.; Esposito, A.; Nicodemi, M.; Matteuzzi, T.; Vercellone, F. Polymer Models of Chromatin Imaging Data in Single Cells. Algorithms 2022, 15, 330. [Google Scholar] [CrossRef]
  153. Trussart, M.; Serra, F.; Baù, D.; Junier, I.; Serrano, L.; Marti-Renom, M.A. Assessing the Limits of Restraint-Based 3D Modeling of Genomes and Genomic Domains. Nucleic Acids Res. 2015, 43, 3465–3477. [Google Scholar] [CrossRef]
  154. Nagano, T.; Lubling, Y.; Stevens, T.J.; Schoenfelder, S.; Yaffe, E.; Dean, W.; Laue, E.D.; Tanay, A.; Fraser, P. Single-Cell Hi-C Reveals Cell-to-Cell Variability in Chromosome Structure. Nature 2013, 502, 59–64. [Google Scholar] [CrossRef]
  155. Stevens, T.J.; Lando, D.; Basu, S.; Atkinson, L.P.; Cao, Y.; Lee, S.F.; Leeb, M.; Wohlfahrt, K.J.; 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. [Google Scholar] [CrossRef]
  156. Demmerle, J.; Innocent, C.; North, A.J.; Ball, G.; Müller, M.; Miron, E.; Matsuda, A.; Dobbie, I.M.; Markaki, Y.; Schermelleh, L. Strategic and Practical Guidelines for Successful Structured Illumination Microscopy. Nat. Protoc. 2017, 12, 988–1010. [Google Scholar] [CrossRef] [PubMed]
  157. Gustafsson, M.G.L.; Shao, L.; Carlton, P.M.; Wang, C.J.R.; Golubovskaya, I.N.; Cande, W.Z.; Agard, D.A.; Sedat, J.W. Three-Dimensional Resolution Doubling in Wide-Field Fluorescence Microscopy by Structured Illumination. Biophys. J. 2008, 94, 4957–4970. [Google Scholar] [CrossRef] [PubMed]
  158. Markaki, Y.; Smeets, D.; Fiedler, S.; Schmid, V.J.; Schermelleh, L.; Cremer, T.; Cremer, M. The Potential of 3D-FISH and Super-resolution Structured Illumination Microscopy for Studies of 3D Nuclear Architecture. BioEssays 2012, 34, 412–426. [Google Scholar] [CrossRef] [PubMed]
  159. Bishop, M.; Michels, J.P.J. Polymer Shapes in Three Dimensions. J. Chem. Phys. 1986, 85, 5961–5962. [Google Scholar] [CrossRef]
  160. Coniglio, A.; Nicodemi, M. A Statistical Mechanics Approach to the Inherent States of Granular Media. Phys. A Stat. Mech. Its Appl. 2001, 296, 451–459. [Google Scholar] [CrossRef]
  161. Nicodemi, M. Force Correlations and Arch Formation in Granular Assemblies. Phys. Rev. Lett. 1998, 80, 1340–1343. [Google Scholar] [CrossRef]
  162. Guha, S.; Mitra, M.K. Multivalent Binding Proteins Can Drive Collapse and Reswelling of Chromatin in Confinement. Soft Matter 2023, 19, 153–163. [Google Scholar] [CrossRef]
  163. Brangwynne, C.P.; Tompa, P.; Pappu, R.V. Polymer Physics of Intracellular Phase Transitions. Nat. Phys. 2015, 11, 899–904. [Google Scholar] [CrossRef]
  164. Choi, J.-M.; Holehouse, A.S.; Pappu, R.V. Physical Principles Underlying the Complex Biology of Intracellular Phase Transitions. Annu. Rev. Biophys. 2020, 49, 107–133. [Google Scholar] [CrossRef]
  165. Scolari, V.F.; Cosentino Lagomarsino, M. Combined Collapse by Bridging and Self-Adhesion in a Prototypical Polymer Model Inspired by the Bacterial Nucleoid. Soft Matter 2015, 11, 1677–1687. [Google Scholar] [CrossRef]
  166. Scolari, V.F.; Mercy, G.; Koszul, R.; Lesne, A.; Mozziconacci, J. Kinetic Signature of Cooperativity in the Irreversible Collapse of a Polymer. Phys. Rev. Lett. 2018, 121, 057801. [Google Scholar] [CrossRef] [PubMed]
  167. Jacobs, W.M.; Frenkel, D. Phase Transitions in Biological Systems with Many Components. Biophys. J. 2017, 112, 683–691. [Google Scholar] [CrossRef] [PubMed]
  168. Ryu, J.-K.; Hwang, D.-E.; Choi, J.-M. Current Understanding of Molecular Phase Separation in Chromosomes. Int. J. Mol. Sci. 2021, 22, 10736. [Google Scholar] [CrossRef] [PubMed]
  169. Ancona, M.; Brackley, C.A. Simulating the Chromatin-Mediated Phase Separation of Model Proteins with Multiple Domains. Biophys. J. 2022, 121, 2600–2612. [Google Scholar] [CrossRef] [PubMed]
  170. Micheletti, C.; Marenduzzo, D.; Orlandini, E. Polymers with Spatial or Topological Constraints: Theoretical and Computational Results. Phys. Rep. 2011, 504, 1–73. [Google Scholar] [CrossRef]
  171. Oliveira, L.P.; Jensen, H.J.; Nicodemi, M.; Sibani, P. Record Dynamics and the Observed Temperature Plateau in the Magnetic Creep-Rate of Type-II Superconductors. Phys. Rev. B—Condens. Matter Mater. Phys. 2005, 71, 104526. [Google Scholar] [CrossRef]
  172. Nicodemi, M.; Fierro, A.; Coniglio, A. Segregation in Hard-Sphere Mixtures under Gravity. An Extension of Edwards Approach with Two Thermodynamical Parameters. Europhys. Lett. 2002, 60, 684–690. [Google Scholar] [CrossRef]
  173. Arenzon, J.J.; Nicodemi, M.; Sellitto, M. Equilibrium Properties of the Ising Frustrated Lattice Gas. J. Phys. I 1996, 6, 1143–1152. [Google Scholar] [CrossRef]
  174. Cataudella, V.; Franzese, G.; Nicodemi, M.; Scala, A.; Coniglio, A. Critical Clusters and Efficient Dynamics for Frustrated Spin Models. Phys. Rev. Lett. 1994, 72, 1541. [Google Scholar] [CrossRef]
  175. Borrelli, A.; De Falco, I.; Della Cioppa, A.; Nicodemi, M.; Trautteur, G. Performance of Genetic Programming to Extract the Trend in Noisy Data Series. Phys. A Stat. Mech. Its Appl. 2006, 370, 104–108. [Google Scholar] [CrossRef]
  176. Caglioti, E.; Coniglio, A.; Herrmann, H.J.; Loreto, V.; Nicodemi, M. Segregation of Granular Mixtures in the Presence of Compaction. Europhys. Lett. 1998, 43, 591. [Google Scholar] [CrossRef]
  177. Tarzia, M.; de Candia, A.; Fierro, A.; Nicodemi, M.; Coniglio, A. Glass Transition in Granular Media. Europhys. Lett. 2004, 66, 531–537. [Google Scholar] [CrossRef]
  178. Nicodemi, M.; Coniglio, A. Macroscopic Glassy Relaxations and Microscopic Motions in a Frustrated Lattice Gas. Phys. Rev. E 1998, 57, R39–R42. [Google Scholar] [CrossRef]
  179. Weischenfeldt, J.; Symmons, O.; Spitz, F.; Korbel, J.O. Phenotypic Impact of Genomic Structural Variation: Insights from and for Human Disease. Nat. Rev. Genet. 2013, 14, 125–138. [Google Scholar] [CrossRef]
  180. Collins, R.L.; Brand, H.; Karczewski, K.J.; Zhao, X.; Alföldi, J.; Francioli, L.C.; Khera, A.V.; Lowther, C.; Gauthier, L.D.; Wang, H.; et al. A Structural Variation Reference for Medical and Population Genetics. Nature 2020, 581, 444–451. [Google Scholar] [CrossRef]
  181. Iannone, F.; Ambrosino, F.; Bracco, G.; De Rosa, M.; Funel, A.; Guarnieri, G.; Migliori, S.; Palombi, F.; Ponti, G.; Santomauro, G.; et al. CRESCO ENEA HPC Clusters: A Working Example of a Multifabric GPFS Spectrum Scale Layout. In Proceedings of the 2019 International Conference on High Performance Computing & Simulation (HPCS), Dublin, Ireland, 15–19 July 2019; IEEE: Piscataway, NJ, USA, 2019; pp. 1051–1052. [Google Scholar]
Figure 1. (a) In the SBS model, a chromatin region is represented by a self-avoiding polymer chain along which specific binding sites are arranged for diffusing cognate molecular binders. By bridging cognate sites on the chain, the binders drive the folding of the polymer-forming microphase-separated globular structures. The SBS binding domains of the studies LTN1 locus (Chr21:28–30 Mb) in human WTC-11 cells are shown along with a schematic cartoon of the polymer model. (b) The polymer gyration radius Rg is shown here as a function of the MD time iteration steps (y-axis normalized by the Rg value at t = 0). The function exhibits a sharp drop around 105 time steps, signaling the collapse of the chain from an initial coil (i.e., randomly folded) to an equilibrium globule conformation [140]. A representative coil and phase-separated globule 3D structures are shown, respectively, below and above the phase transition point (gray shaded line in the figure).
Figure 1. (a) In the SBS model, a chromatin region is represented by a self-avoiding polymer chain along which specific binding sites are arranged for diffusing cognate molecular binders. By bridging cognate sites on the chain, the binders drive the folding of the polymer-forming microphase-separated globular structures. The SBS binding domains of the studies LTN1 locus (Chr21:28–30 Mb) in human WTC-11 cells are shown along with a schematic cartoon of the polymer model. (b) The polymer gyration radius Rg is shown here as a function of the MD time iteration steps (y-axis normalized by the Rg value at t = 0). The function exhibits a sharp drop around 105 time steps, signaling the collapse of the chain from an initial coil (i.e., randomly folded) to an equilibrium globule conformation [140]. A representative coil and phase-separated globule 3D structures are shown, respectively, below and above the phase transition point (gray shaded line in the figure).
Ijms 25 10215 g001
Figure 2. The gyration radius of the polymer chain is shown here as a function of (a) the binder concentration, c (expressed in volume fraction), and (b) the bead-binder affinity (Ebb, in KBT units). For c ≃ 0 (i.e., no binders), or analogously for Ebb ≃ 0 (i.e., no bead-binder attractions), the polymer is a randomly folded chain (in the SAW universality class), as only random and fleeting contacts are established in the absence of binders [140]; as soon as the number of binders (or their energy affinity) grows above a characteristic threshold, the polymer collapses into an equilibrium globular state where specific contacts are established based on the underlying distribution of the inferred binding sites (see above). The characteristic threshold concentrations and affinities depend on model details: for the considered model parameters, they fall, respectively, around 0.1 and 1.0 KBT.
Figure 2. The gyration radius of the polymer chain is shown here as a function of (a) the binder concentration, c (expressed in volume fraction), and (b) the bead-binder affinity (Ebb, in KBT units). For c ≃ 0 (i.e., no binders), or analogously for Ebb ≃ 0 (i.e., no bead-binder attractions), the polymer is a randomly folded chain (in the SAW universality class), as only random and fleeting contacts are established in the absence of binders [140]; as soon as the number of binders (or their energy affinity) grows above a characteristic threshold, the polymer collapses into an equilibrium globular state where specific contacts are established based on the underlying distribution of the inferred binding sites (see above). The characteristic threshold concentrations and affinities depend on model details: for the considered model parameters, they fall, respectively, around 0.1 and 1.0 KBT.
Ijms 25 10215 g002
Figure 3. (a) In situ Hi-C 2.0 contact data of the studied 2 Mb wide LTN1 locus in WTC-11 cells (left) are consistently captured by the SBS polymer model (right). The high Pearson and genomic distance corrected correlation values (respectively, r = 0.90 and r’ = 0.59) indicate that the model accurately captures the overall structure of LTN1 pairwise interactions. (b) Comparison between the model and Hi-C contact probabilities at the LTN1 locus in WTC-11 cells. Overall, the model consistently recapitulates the experimental profile across genomic scales (r = 0.98).
Figure 3. (a) In situ Hi-C 2.0 contact data of the studied 2 Mb wide LTN1 locus in WTC-11 cells (left) are consistently captured by the SBS polymer model (right). The high Pearson and genomic distance corrected correlation values (respectively, r = 0.90 and r’ = 0.59) indicate that the model accurately captures the overall structure of LTN1 pairwise interactions. (b) Comparison between the model and Hi-C contact probabilities at the LTN1 locus in WTC-11 cells. Overall, the model consistently recapitulates the experimental profile across genomic scales (r = 0.98).
Ijms 25 10215 g003
Figure 4. (a) Representative examples of model-predicted phase-separated single-molecule distance matrices of the LTN1 locus. The interaction patterns broadly differ across the ensemble of polymer configurations, as the system can fold in a variety of 3D architectures [127]. (b) The structural heterogeneity of individual polymer structures is measured in the model by computing the r’ correlation between pairs of single-molecule distance matrices. The resulting distribution is broad (variance = 0.15) and has a non-zero average value (mean = 0.13), indicating that chromatin structures are highly variable from cell to cell yet have a residual degree of structural correlation.
Figure 4. (a) Representative examples of model-predicted phase-separated single-molecule distance matrices of the LTN1 locus. The interaction patterns broadly differ across the ensemble of polymer configurations, as the system can fold in a variety of 3D architectures [127]. (b) The structural heterogeneity of individual polymer structures is measured in the model by computing the r’ correlation between pairs of single-molecule distance matrices. The resulting distribution is broad (variance = 0.15) and has a non-zero average value (mean = 0.13), indicating that chromatin structures are highly variable from cell to cell yet have a residual degree of structural correlation.
Ijms 25 10215 g004
Figure 5. (a) Distribution of ellipsoid semi-axis ratios, a/c (histogram in blue) and b/c (in orange), computed from the inertia tensor of single-molecule polymer structures. The black dashed line in the figure represents the expected value in the case of perfectly spherical conformations (a = b = c). Interestingly, model-predicted 3D structures of the LTN1 locus appear prolate, as we find a > bc. The two distributions, a/c and b/c, are statistically different from each other (two-sided Mann–Whitney p-value < 0.001). (b) The ellipticity of model single molecules, calculated from the eigenvalues of their gyration tensor [159], exhibits a broad distribution (variance = 0.16) with an average value of 0.51, indicating a significant structural variability and a tendency towards a prolate shape.
Figure 5. (a) Distribution of ellipsoid semi-axis ratios, a/c (histogram in blue) and b/c (in orange), computed from the inertia tensor of single-molecule polymer structures. The black dashed line in the figure represents the expected value in the case of perfectly spherical conformations (a = b = c). Interestingly, model-predicted 3D structures of the LTN1 locus appear prolate, as we find a > bc. The two distributions, a/c and b/c, are statistically different from each other (two-sided Mann–Whitney p-value < 0.001). (b) The ellipticity of model single molecules, calculated from the eigenvalues of their gyration tensor [159], exhibits a broad distribution (variance = 0.16) with an average value of 0.51, indicating a significant structural variability and a tendency towards a prolate shape.
Ijms 25 10215 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Conte, M.; Abraham, A.; Esposito, A.; Yang, L.; Gibcus, J.H.; Parsi, K.M.; Vercellone, F.; Fontana, A.; Di Pierno, F.; Dekker, J.; et al. Polymer Physics Models Reveal Structural Folding Features of Single-Molecule Gene Chromatin Conformations. Int. J. Mol. Sci. 2024, 25, 10215. https://doi.org/10.3390/ijms251810215

AMA Style

Conte M, Abraham A, Esposito A, Yang L, Gibcus JH, Parsi KM, Vercellone F, Fontana A, Di Pierno F, Dekker J, et al. Polymer Physics Models Reveal Structural Folding Features of Single-Molecule Gene Chromatin Conformations. International Journal of Molecular Sciences. 2024; 25(18):10215. https://doi.org/10.3390/ijms251810215

Chicago/Turabian Style

Conte, Mattia, Alex Abraham, Andrea Esposito, Liyan Yang, Johan H. Gibcus, Krishna M. Parsi, Francesca Vercellone, Andrea Fontana, Florinda Di Pierno, Job Dekker, and et al. 2024. "Polymer Physics Models Reveal Structural Folding Features of Single-Molecule Gene Chromatin Conformations" International Journal of Molecular Sciences 25, no. 18: 10215. https://doi.org/10.3390/ijms251810215

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop